You are on page 1of 35

Signal Transduction and Targeted Therapy www.nature.

com/sigtrans

REVIEW ARTICLE OPEN

Targeting cancer stem cell pathways for cancer therapy


Liqun Yang1,2, Pengfei Shi1,2, Gaichao Zhao1,2, Jie Xu1,2, Wen Peng1,2, Jiayi Zhang1,2, Guanghui Zhang1,2, Xiaowen Wang1,2,
Zhen Dong1,2, Fei Chen3 and Hongjuan Cui1,2

Since cancer stem cells (CSCs) were first identified in leukemia in 1994, they have been considered promising therapeutic targets for
cancer therapy. These cells have self-renewal capacity and differentiation potential and contribute to multiple tumor malignancies,
such as recurrence, metastasis, heterogeneity, multidrug resistance, and radiation resistance. The biological activities of CSCs are
regulated by several pluripotent transcription factors, such as OCT4, Sox2, Nanog, KLF4, and MYC. In addition, many intracellular
signaling pathways, such as Wnt, NF-κB (nuclear factor-κB), Notch, Hedgehog, JAK-STAT (Janus kinase/signal transducers and
activators of transcription), PI3K/AKT/mTOR (phosphoinositide 3-kinase/AKT/mammalian target of rapamycin), TGF (transforming
growth factor)/SMAD, and PPAR (peroxisome proliferator-activated receptor), as well as extracellular factors, such as vascular niches,
hypoxia, tumor-associated macrophages, cancer-associated fibroblasts, cancer-associated mesenchymal stem cells, extracellular
matrix, and exosomes, have been shown to be very important regulators of CSCs. Molecules, vaccines, antibodies, and CAR-T
(chimeric antigen receptor T cell) cells have been developed to specifically target CSCs, and some of these factors are already
undergoing clinical trials. This review summarizes the characterization and identification of CSCs, depicts major factors and
pathways that regulate CSC development, and discusses potential targeted therapy for CSCs.
1234567890();,:

Signal Transduction and Targeted Therapy (2020)5:8 ; https://doi.org/10.1038/s41392-020-0110-5

INTRODUCTION targeting CSCs for cancer therapy both in basic research and
Cancers are chronologic diseases that seriously threaten human clinical studies.
life. Many strategies have been developed for cancer treatment,
including surgery, radiotherapy, chemotherapy, and targeted
therapy. Because of all these treatments, the incidence rate of THE CONCEPT OF CSCS
cancer has been stable in women and has declined slightly in Biological characteristics of CSCs
men in the past decade (2006–2015), and the cancer death rate With the deepening of tumor biology research, clinical diagnosis
(2007–2016) also declined.1 However, traditional cancer treat- and cancer treatment have significantly improved in recent years.
ment methods are effective only for some malignant tumors.2 However, the high recurrence rate and high mortality rate are still
The main reasons for the failure of cancer treatment are unresolved and are closely related to the biological characteristics
metastasis, recurrence, heterogeneity, resistance to chemother- of CSCs. With further understanding of CSC characteristics,
apy and radiotherapy, and avoidance of immunological surveil- research on tumor biology has entered a new era. Therefore,
lance.3 All these failures could be explained by the characteristics understanding the biological properties of CSCs is of great
of cancer stem cells (CSCs).4 CSCs can cause cancer relapse, significance in the diagnosis and treatment of tumors.
metastasis, multidrug resistance, and radiation resistance through CSCs have a strong self-renewal ability, which is the direct cause
their ability to arrest in the G0 phase, giving rise to new tumors.5 of tumorigenesis.12 CSCs can symmetrically divide into two CSCs
Therefore, CSCs could be considered the most promising targets or into one CSC and one daughter cell.13 CSCs expand in a
for cancer treatment. symmetrical splitting manner to excessively increase cell growth,
CSCs were first identified in leukemia and then isolated via ultimately leading to tumor formation.14 CSCs isolated from
CD34+ and CD38− surface marker expression in the 1990s.6,7 CSCs original tumor tissue that were transplanted into severe combined
expressing different surface markers, such as CD133, nestin, and immunodeficiency disease (SCID) mice then formed new tumors.15
CD44, have been subsequently found in many nonsolid and solid CSCs and normal stem cells also share some of the same
tumors, and these cells also form the bulk of the tumor.8,9 CSCs regulatory signaling pathways, such as the Wnt/β-catenin,16 Sonic
can generate tumors via the self-renewal and differentiation into Hedgehog (Hh),17 and Notch pathways, which are involved in the
multiple cellular subtypes.10 The activities of CSCs are controlled self-renewal process.18 In addition, other signaling molecules, such
by many intracellular and extracellular factors, and these factors as PTEN and the polycomb family, also play important roles in the
can be used as drug targets for cancer treatment.11 To understand regulation of CSC growth.19 The regulation of CSC self-renewal is
the nature of CSCs, we summarized their characteristics, methods the key link to understanding tumorigenesis. These studies will
for identification and isolation, regulation and current research on provide a clear target for cancer treatment.

1
State Key Laboratory of Silkworm Genome Biology, Southwest University, 400716 Chongqing, China; 2Cancer Center, Medical Research Institute, Southwest University, 400716
Chongqing, China and 3Department of Pharmaceutical Sciences, Eugene Applebaum College of Pharmacy and Health Sciences, Wayne State University, Detroit, MI 48201, USA
Correspondence: Hongjuan Cui (hongjuan.cui@gmail.com)
These authors contributed equally: Liqun Yang, Pengfei Shi

Received: 5 October 2019 Revised: 15 December 2019 Accepted: 19 December 2019

© The Author(s) 2020


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
2
In addition to their self-renewal ability, CSCs also have the challenging to isolate and identify CSCs. However, an increasing
ability to differentiate into different cell types. Bonnet and Dick7 number of techniques and means have emerged.
demonstrated in 1997 that CD34+/CD38− leukemia stem cells CSCs have been identified through different biomarkers in
(LSCs) have the ability to differentiate and proliferate in SCID mice. human cancers (Table 1). CSCs can be separated by combining
Brain CSCs isolated from patients are positive for the markers specific biomarkers that are mostly located on the cell surface.3
CD133 and nestin, which are the same markers as those of normal The primary separation techniques are fluorescence-activated cell
neuronal stem cells, but some cells lack surface markers for sorting (FACS) and magnetic-activated cell sorting (MACS).44,45
differentiation.20 Generally, various signaling pathways regulate Since Dick JE first screened CSCs from leukemia by using FACS
the self-renewal and differentiation of normal stem cells to technology,7 FACS has become the most widely used technique
promote their proliferation and differentiation in a relatively for cell separation. It can perform multibiomarker sorting at one
balanced manner. Once the regulatory balance is destroyed, time and has high purity and strong specificity. MACS is a MACS
uncontrolled CSCs ultimately lead to tumorigenesis.21 CSCs also technique. MACS separation is relatively simple, but the technique
transdifferentiate into other multilineage cells to regulate is cumbersome. Therefore, this method requires high activity of
tumorigenesis.22 Bussolati et al.23 found that renal CSCs differ- CSCs.44,46 These two methods are effective in separating CSCs
entiated into vascular endothelial cells (ECs) in the bulk of tumors from large numbers of cells.
formed in SCID mice after injection of human renal CSCs. Additionally, there are other ways to separate CSCs from
Additionally, CSCs that differentiate into vascular ECs and promote tumors. In 1996, Dr. Goodell observed that after adding Hoechst
angiogenesis have been found in a variety of cancers, such as 33342 to a culture of bone marrow cells, a few cells did not
glioblastoma24 and liver cancer.25 accumulate dyes, and he claimed that these few cells were side
Metastasis refers to the process by which cancer cells travel population (SP) cells. Therefore, SP cells can be separated by
from the primary site through lymphatic vessels, blood vessels, or fluorescence screening after the outflow of Hoechst 33342.
the body cavity.26 Since stromal cells (such as granulocytes and Recently, SP cells have been identified in various normal tissues
macrophages) secrete signaling molecules in the tumor micro- and tumor cells. SP cells have high homology, self-renewal and
environment (TME), these cells stimulate epithelial–mesenchymal multidirectional differentiation potential.47,48 Some reports have
transformation (EMT) to promote the invasion of tumor cells,27 shown that ABCG2 is highly expressed in SP cells.47,49 ABCG2 is
which induce differentiated human mammary epithelial cells to highly related to the drug resistance of CSCs and is used as a
form mammary glands.28 Activation of the RAS/MAPK (mitogen- phenotypic marker for CSCs,50,51 including ovarian cancer,52
activated protein kinase) signaling pathway transforms nontu- AML,53 breast cancer,54 lung cancer,55 nasopharyngeal carci-
morigenic CD44−/CD24+ breast cancer cells into tumorigenic noma,56 and hepatocellular carcinoma (HCC).57 Montanaro
CD44+/CD24− breast cancer cells.29 A study showed that CSCs are et al.58 explored the optimal concentration of Hoechst 33342 to
closely related to EMT, and EMT is likely to be the basis for tumor reduce the toxic effect. The SP sorting method has universal
invasion and metastasis. In addition, CD133+/CXCR4+ pancreatic applicability in the separation and identification of CSCs, especially
cancer cells30 and CD44+/α2βhi1/CD133+ prostate cancer cells31 CSCs with unknown cell surface markers, and is an effective
are also tumorigenic. Therefore, these studies indicate that CSCs method for CSC research.
play a crucial role in tumor metastasis and development. The colony-forming ability of CSCs is also used for separation
Furthermore, understanding the mechanism of CSC drug and identification.59 After digestion of the tumor tissues into
resistance is vital for cancer treatment and preventing recur- single cells, low-density cell culture can be conducted in serum-
rence.32 CSCs efficiently express ATP-binding cassette (ABC) free medium containing epithelial growth factor (EGF) and basic
transporters (including MDR1 (ABCB1), MRP1 (ABCC1), and fibroblast growth factor (FGF).60 Under this condition, a single CSC
(ABCG2)), which are multidrug resistance proteins, and these will form a cell colony or sphere. Taylor et al.61 successfully
proteins protect leukemia and some solid tumor cells from drug isolated CSCs from a variety of neurological tumors by using this
damage and induce drug resistance.33 According to previous colony formation assay. However, the cell purification rate is low,
studies, aldehyde dehydrogenase (ALDH), a marker in many and the CSC specificity is poor in this assay. The in vivo limited
CSCs,34 eliminates oxidative stress and enhances resistance to dilution assay (LDA) can be used for assessing CSC activity. After
chemotherapeutic drugs, such as oxazolidine, taxanes, and low-density transplantation of immune-deficient mice with the
platinum drugs.35 ALDH also removes free radicals induced by limiting dilution method, CSCs can be identified by ELDA software
radiation and stimulates resistance to radiation.35 Inducing DNA analysis, and this method is affected by cell density and the
damage and apoptosis through chemotherapy and radiotherapy microenvironment in mice.62
are commonly used cancer treatments. However, CSCs can Traditional chemotherapeutic drugs mainly affect cancer cells,
effectively protect cancer cells from apoptosis by activating DNA but CSCs are mostly arrested in the G0 phase and are relatively
repair abilities.36 static, thus evading the killing effect of chemotherapeutic drugs.63
It is currently believed that CSCs are the key "seeds" for tumor Hence, the drug-resistant characteristics of CSCs can be used to
initiation and development, metastasis, and recurrence.37 CSCs isolate and identify CSCs.64 Previous studies have shown that
have evolved and are highly heterogeneous.38 Breast CSCs have radiotherapy combined with hypoxic culture can also be used to
different expression patterns of surface biomarkers, such as enrich CSCs.65 In addition, the separation of CSCs can also be
CD44+, CD24−, SP, and ALDH+.29,34,39 CD271− or CD271+ accomplished by physical methods. Hepatoma stem cells can be
melanoma stem cells can form tumors in SCID mice.40 The isolated from rat liver cancer tissue by Percoll density gradient
heterogeneity of CSCs has also been found in other cancers, centrifugation; a cell fraction with a high nuclear-to-cytoplasmic
including glioblastoma,41 prostate cancer,42 and lung cancer.43 ratio is obtained.66 Recently, Rahimi et al.67 used the miR-302 host
The heterogeneity of CSCs is so complex that more effective gene promoter to overexpress neomycin in cancer cells and
biomarkers are needed to identify CSCs or distinguish the selected and collected neomycin-resistant CSCs.
heterogeneity of CSCs.

Isolation and identification of CSCs FACTORS REGULATING CSCS


It is known that the proportion of CSCs in tumor tissues is very low CSCs can originate from at least four cell types, including normal
and generally accounts for only 0.01–2% of the total tumor mass. stem cells, directed group progenitor cells, mature cells, and the
In addition, CSCs and normal stem cells also share similar fusion of stem cells and other mutant cells.68 Therefore,
transcription factors and signaling pathways. Therefore, it is more transformed CSCs from normal cells require multiple gene

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
3
Table 1. Various biomarkers of cancer stem cells in human cancers

Cancers Markers Function

Breast CD29+658, ALDH: An enzyme that plays a role in cell resistance665


CD49f+659, CD44: A glycoprotein involves in cell migration and self-renewal666
CD90+660, CD90: A glycoprotein participates in T cell adhesion and signal
CD133+661, transduction667
ALDH+662, CD133: A transmembrane glycoprotein that maintains lipid composition in
ESA+/CD44+/CD24,663 cell membranes668
CD44+/CD24−664 CD24: A marker that promotes blood flow in the tumor during metastasis669
CD49f: A membrane proteins of the integrin family that plays an important
role in cell surface adhesion and signaling670
Prostate EpCAM+671, α2β1: A receptor involves in cell adhesion and recognition31
CD117+672, E-cadherin: It plays an important role in tumor migration and invasion676
α2β1+31, CXCR4: CXC chemokine receptor works with CD4 protein to support HIV
ALDH+42, entry into cells675
CD44+673, EZH2: A member of the Polycomb family plays an vital role in the central
EZH2+674, nervous system674
CXCR4+675,
E-cadherin+676,
CD133+677
Brain CD49f+678, CD36: The main glycoprotein on the surface of platelet has an important
CD90+679, function as an adhesion molecule686
CD44+680, EGFR: It binds to epidermal growth factor and promote proliferative
CD36+681, migration in tumors682
EGFR+682, A2B5: A ganglioside marker that identifies subpopulations of nerve cells in
A2B5+683, the central nervous system687
L1CAM+684, L1CAM: A adhesion molecule that plays an important role in the
CD133+41,685 development of the nervous system include neuronal migration and
differentiation684
Stomach ALDH+688, CD44+689, CD44V8–10: A variant of CD44 with a specific class of CSCs690
CD44V8–10+690, CD54: A class of adhesion molecules express in malignant tumor cells693
CD133+691, CD24+692,
CD54+693, CD90+694,
CD49f+678 CD71+695,
EpCAM+696
Colorectal CD200+697, EpCAM+698, CD200: A glycoprotein plays an important role in the regulation of
CD133+699, CD166+, immunosuppression and anti-tumor activity703
CD206+700, CD44+701, CD166: It binds to the T cell differentiation antigen CD6 and involves in cell
CD49f+678, ALDH+702 adhesion and migration processes704
CD206: A mannose receptor involves in endocytosis, phagocytosis, and
immune homeostasis700
EpCAM: It expresses on most normal epithelial cells and gastrointestinal
cancers, and acts as a homotypic calcium-independent cell adhesion
molecule705
Liver CD24+706, CD133+707, CD13: A receptor for human coronavirus strains, which is the main cause of
CD13+708, CD44+709, upper respiratory tract infection and leukemia712
CD206+700, OV-6+708, OV-6: A marker for rat oval cells and hepatic stem cells708
CD90+710, EpCAM+711
AML CD34+, CD34: It plays a role in the attachment of stem cells to bone marrow
CD38−, extracellular or stromal cells713
CD90+, CD38: An intracellular Ca2+ mobilization messenger, prognostic markers for
CD71+, patients with chronic lymphocytic leukemia714
CD19+, CD71: A transferrin receptor is important for nerve development715
CD20+, CD19: A class of signal transduction molecules regulate B lymphocyte
CD44+, differentiation716
CD10+, CD20: The protein plays a role in the development and differentiation of B
CD45RA+, cells into plasma cells717
CD123+15 CD10: It inhibits a variety of peptide hormones, include glucagon,
encephalin, oxytocin, and bradykinin718
CD45RA: A class of leukocyte activation regulators719
CD123: An interleukin-specific subunit of a heterodimeric cytokine
receptor720
Melanoma CD20+721, CD271+,722, CD271: A nerve growth factor receptor mediates cell survival and cell death
ALDH+723, CD133+724 in nerve cells725
Bladder CD44v6+726, CD44+727, CD44v6: It involves in cell migration, cell adhesion729
ALDH+728
Ovarian CD24+730, ALDH+,731, CD117: A class of transmembrane receptors is also known as stem cell
CD44+/CD117+732, factors735
EpCAM+733, CD133+734

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
4
Table 1 continued
Cancers Markers Function

Pancreas ALDH+736, CD133+30, ABCG2: A class of membrane proteins belongs to the ABC transporter
CD44+/CD24+/EpCAM+17, superfamily that plays a role in the drug resistance properties of CSCs
ABCG2+737, CXCR4+,738
HNSCC ALDH+739, CD44+,740,
CD166+741
Gallbladder CD44+/CD133+742
RCC CD133+743, ALDH+,743, CD105: TGF receptor that involves in TGF-β signaling plays a role in
CXCR4+743, CD44+,744, angiogenesis745
CD105+23
Lung CD166+746, CD90+,747, CD87: A receptor for urokinase plasminogen activator that affects many
CD87+748, ALDH+,749, normal and pathological processes associates with cell surface
CD44+750, CD133+751 plasminogen activation and local degradation of extracellular matrices748
Malignant CD9+, CD9: A glycoprotein plays a role in many cellular processes, includes
mesothelioma CD24+, differentiation, adhesion and signal transduction, and plays a key role in
CD26+752 cancer cell movement and metastasis753
CD26: A class of serine exopeptidases is also an intrinsic membrane
glycoprotein754
OSCC CD44+/CD24,−755 ITGA7: A integrin plays a role in cell migration, morphogenesis,
ITGA7+756 differentiation, and metastasis and participates in the process of
differentiation and migration during myogenesis757
cSCC CD44+758, CD133+759
Esophageal ITGA7+, CD44+,
ALDH+, CD133+,
CD90+297
MM CD138−, CD138: A member of the Syndecan proteoglycan family that involves in cell
CD19+, proliferation, cell migration, and cell–matrix interactions762
CD27+760,761 CD27: A transmembrane glycoprotein involves in the regulation of B cell
activation and immunoglobulin synthesis763
Cervix ABCG2+, CD133+, CD49f+764, ALDH+765
Nasopharyngeal CD44+766, CD133+767,
ALDH+768, CD24+769
Laryngeal ALDH+, CD44+770, CD133+771
AML acute myeloid leukemia, HNSCC head and neck squamous cell carcinoma, RCC renal cell carcinoma, OSCC oral squamous cell carcinoma, cSCC cutaneous
squamous cell carcinoma, MM multiple myeloma, ALDH aldehyde dehydrogenase, EpCAM epithelial cellular adhesion molecule

mutations, epigenetic changes, uncontrolled signaling path- factors.73 Recently, Oct4 has emerged as a master regulator that
ways, and continuous regulation of the microenvironment. It is controls pluripotency, self-renewal, and maintenance of stem
currently believed that there are many similarities between CSCs cells.74 Some studies have reported that Oct4 is highly
and embryonic stem (ES) cells, especially regarding their ability expressed in CSCs.70,73 High expression of Oct4 is positively
to grow indefinitely and self-renew, signaling pathways and correlated with glioma grades75 and promotes self-renewal,
some transcription factors. In addition, CSCs exist in the chemoresistance, and tumorigenicity of HCC stem cells.76 High
supporting microenvironment, which is vital for their survival. expression of Oct4 is also observed in breast CSC-like cells
Moreover, the complex interaction between CSCs and their (CD44+/CD24−).77 Cisplatin, etoposide, adriamycin, and pacli-
microenvironment can further regulate CSC growth. This section taxel γ-irradiation upregulate the expression of Oct4 in lung
will discuss the effects of transcription factors, signaling path- cancer cells, and CD133+ cells are more resistant to drug
ways, and the microenvironment on CSC survival, apoptosis, and treatments than CD133− cells.78 Data also show that Oct4
metastasis. expression is associated with poor clinical outcome in hormone
receptor-positive breast cancer.79 Knockdown of Oct4 also
Major transcription factors in CSCs reduces the stemness of germ cell tumors.80 Hence, these
Generally, stem cells have at least two common characteristics: the studies have proven that Oct4 is a pluripotent factor in CSCs.
ability to self-renew and the potential to differentiate into one or Sox2 belongs to the family of high-mobility group transcription
more specialized cell types.69 Somatic cells can be reprogrammed factors and plays a significant function in the early development
to become induced pluripotent stem cells by transient ectopic and maintenance of undifferentiated ESCs. It is also one of the key
overexpression of the transcription factors Oct4, Sox2, Nanog, transcription factors in CSCs. Rodriguez-Pinilla et al.81 found that
KLF4, and MYC.70–72 In addition, there are some similarities increased expression of Sox2 in basal-like breast cancer may help
between CSCs and ES cells. It is reasonable that some embryonic to characterize poorly differentiated/stem cell phenotypes.82
transcription factors can be re-expressed or reactivated in CSCs.69 Hagerstrand et al.82 also found that a high level of Sox2 can
Therefore, these transcription factors play a very important role in induce xenograft glioma. Further studies showed that knockout of
the regulation of CSC growth. Sox2 inhibits glioblastoma cell proliferation and tumorigenicity,
Oct4, a homeodomain transcription factor of the Pit-Oct-Unc which suggests that Sox2 is the basis for maintaining the self-
family, is recognized as one of the most important transcription renewal ability of tumor-initiating cells (TICs).83 Sox2 also

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
5
maintains the self-renewal of TICs in osteosarcomas, and down- stagnation of proliferation, aging, or apoptosis.110 MYC is usually
regulation of Sox2 drastically decreases its transformative deregulated in human cancers, plays an important role in
characteristics and tumorigenesis ability in vitro. Furthermore, maintaining the number of invasive CSCs,111 and is also one of
osteosarcoma cells that lose Sox2 cannot form osteospheres and the most effective oncogenes for detecting the cell transformation
differentiate into mature osteoblasts any longer.84 Sox2 is found in phenotype in vitro and in vivo. Previous studies have shown that
invasive cutaneous squamous cell carcinoma (SCC) and promotes deletion of the tumor suppressor gene p53 and MYC synergizes to
the metastasis of cancer cells.85 These studies suggest that Sox2 induce hepatocyte proliferation and tumorigenesis.112 In addition
promotes self-renewal and tumorigenesis and inhibits differentia- to p53 deletion, overexpression of Bcl-2 and Bmi-1 and loss of
tion in CSCs. p19ARF also assist MYC in regulating the survival and proliferation
Nanog, a differentiated homeobox (HOX) domain protein that of CSCs.113 The expression of the three members of the MYC
was first discovered in ESCs, has typical self-renewal and multi- family is different in different tumors, such as C-MYC in leukemia
potent transcriptional regulatory functions.86 Although Nanog is and tongue SCC stem cells114,115 and N-MYC in small-cell lung
silenced in normal somatic cells, abnormal expression has been cancer, prostate cancer, neuroblastoma, and medulloblas-
reported in human cancers, such as breast cancer, cervical cancer, toma.116,117 L-MYC is expressed in hematopoietic malignancies.118
brain cancer, colon cancer, head and neck cancer, lung cancer, In addition, inactivation of MYC results in HCC stem cells
and gastric cancer.86–90 Compared to levels in benign tissues, differentiating into hepatocytes and biliary duct cells to form bile
Nanog messenger RNA (mRNA) is elevated in malignant tumors. In duct structures, which might be associated with the loss of the
a number of patients with colorectal cancer (n = 175), high Nanog tumor marker α-fetoprotein and increased expression of cytoker-
protein is associated with lymph node positivity and Dukes atin 8, hepatocyte markers, carcinoembryonic antigen, and the
grade.91 Similarly, overexpression of Nanog in colorectal CSCs liver stem cell marker cytokeratin 19.119 Studies have also shown
promotes colony formation and tumorigenicity in vivo.92 In that MYC is highly expressed in glioblastoma multiforme stem
addition, gastric cancer patients with high Nanog levels have a cells and induces cell proliferation and invasion and inhibits
lower 5-year survival rate.88 The expression level of Nanog is apoptosis.111 Increased copy number of the MYC gene in human
increased in HCC cell lines and primary tumors and is associated and mouse prostate CSCs has also been found.120 These studies
with advanced diseases (tumor node metastasis (TNM) stage III/ indicate that MYC induces tumorigenesis with the help of other
IV).93 Through the study of prostatic cell lines, xenografts and factors.
primary tumors, it was found that Nanog short hairpin RNA
inhibits the formation of primary prostate cancer cells (PCA) Major signaling pathways in CSCs
spheres, clonal growth, and tumorigenesis.94 In 43 cases of Many signaling pathways that contribute to the survival,
pancreatic cancer tissue microarray analysis, Kaplan–Meier analy- proliferation, self-renewal, and differentiation properties of normal
sis showed that high expression of Nanog (and Oct4) predicted stem cells are abnormally activated or repressed in tumorigenesis
worse prognosis and was negatively correlated with patient or CSCs. Many endogenous or exogenous genes and microRNAs
survival.95 These studies indicate that Nanog plays an important regulate these complex pathways. These signaling pathways can
role in regulating the self-renewal and proliferation of CSCs. also induce downstream gene expression, such as cytokines,
KLF4 is expressed in many tissues and plays an important role in growth factors, apoptosis genes, antiapoptotic genes, proliferation
many different physiological processes. As a bifunctional tran- genes, and metastasis genes in CSCs. These signaling pathways
scription factor, KLF4 activates or inhibits transcription according are not a single regulator but interwoven networks of signaling
to different target genes and utilizing different mechanisms. KLF4 mediators to regulate CSC growth. Therefore, this section will
can play an oncogenic or anticancer role, depending on the type describe how signaling pathways regulate CSC growth.
of cancer involved. For example, KLF4 is an anticancer factor in the
intestinal epithelium and gastric epithelium.96 The expression of Wnt signaling pathway in CSCs. Wnts include large protein
KLF4 is downregulated with hypermethylation and loss of ligands that affect diverse processes, such as the generation of cell
heterozygosity in colorectal CSCs and gastric CSCs.97 Down- polarity, and cells fate.121 The Wnt pathway is highly complex and
regulation of KLF4 is also found in other cancers, such as non- evolutionarily conserved and includes 19 Wnt ligands and more
small-cell lung carcinoma,98 liver cancer,99 leukemia,100 anaplastic than 15 receptors.122 The Wnt signaling pathway can be divided
meningioma,101 bladder cancer,102 and esophageal cancer.103 into canonical Wnt signaling (through the FZD-LRP5/6 receptor
Although these data clearly demonstrate that KLF4 plays an complex, leading to derepression of β-catenin) and noncanonical
anticancer role in those cancers, KLF4 may also be an oncogene, Wnt signaling (through FZD receptors and/or ROR1/ROR2/RYK
which was demonstrated for the first time in nearly a decade.104 coreceptors, activating PCP, RTK, or Ca2+ signaling cascades).123 In
Overexpression of KLF4 in transformed rat renal epithelial cells canonical Wnt signaling, in the absence of Wnt ligands (inactive
induces tumorigenesis of laryngeal SCC.105 In addition, depletion Wnt signaling state, Fig. 1, left), β-catenin is phosphorylated by
of KLF4 inhibits melanoma xenograft growth in vivo.106 High glycogen synthase kinase 3β (GSK3β), which leads to β-catenin
expression of KLF4, an oncogene in human breast CSCs, is degradation via β-TrCP200 ubiquitination and inhibits transloca-
correlated with an aggressive phenotype in canine mammary tion of β-catenin from the cytoplasm to the nucleus.124 In contrast,
tumors.107 These studies suggest that KLF4 has different functions in the presence of Wnt ligands (e.g., Wnt3a and Wnt1), the ligands
in different CSCs. combine with Fzd receptors and LRP coreceptors (active Wnt
MYC has three family members (C-Myc, N-Myc, and L-Myc, signaling, Fig. 1, right). LRP receptors are phosphorylated by
which are encoded by the proto-oncogene family and are GSK3β and CK1α.125 β-Catenin is released from the Axin complex
essential transcription factors in the DNA-binding proteins of the to enter the nucleus. In addition, β-catenin combines with LEF/TCF
basic helix–loop–helix (bHLH) superfamily). MYC regulates a large and enhances the recruitment of histone-modifying coactivators,
number of protein-coding and noncoding genes and coordinates such as BCL9, Pygo, CBP/p300, and BRG1, to activate transcription.
various biological processes in stem cells, such as cell metabolism, Noncanonical Wnt signaling does not involve β-catenin. During
self-renewal, differentiation, and growth.108,109 Although the MYC Wnt/PCP signaling, Dvl is activated through binding of Wnt
gene is one of the most commonly activated oncogenes that is ligands and the ROR-Frizzled receptor.126 Dvl inhibits the binding
involved in the pathogenesis of human cancer, overexpression of of the small GTPase Rho and the cytoplasmic protein DAAM1.127
MYC alone is surprisingly unable to induce the transformation of The small GTPases Rac1 and Rho together trigger ROCK (Rho
normal cells into tumor cells. The overexpression of MYC in normal kinase) and JNK (c-Jun N-terminal kinase). This results in
human cells may be ineffective or highly destructive, resulting in cytoskeletal rearrangement and/or transcriptional responses.128

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
6

Fig. 1 Wnt/β-catenin pathway in cancer stem cells. The canonical Wnt/β-catenin pathway regulates the pluripotency of CSCs and determines
the differentiation fate of CSCs. In the absence of Wnt signaling, β-catenin is bound to the Axin complex, which contains APC and GSK3β, and
is phosphorylated, leading to ubiquitination and proteasomal degradation through the β-Trcp pathway. However, the complex (TAZ/YAP), the
long noncoding RNA TIC1 and proteins (TRAP1 and TIAM1) regulate the β-Trcp pathway. In the presence of Wnt signaling, the binding of
LRP5/6 and Fzd inhibits the activity of the Axin complex and the phosphorylation of β-catenin, which makes β-catenin enter the nucleus, and
then bind to TEF/TCF to form a complex, which then recruits cofactors to initiate downstream gene expression. Some proteins (DKK2
(Dickkopf-related protein 2), DACT1, CDH11, GECG, PKM2, EZH2, CD44v6, MYC, and TERT), microRNAs (miR-1246, miR-9, miR-92a, miR-544a,
and miR-483-5p), and long noncoding RNAs (lncR-β-catm and lncR-TCF7) regulate the activation of the Wnt/β-catenin pathway in CSCs

Wnt/Ca2+ signaling is activated by G protein-triggered phospho- diffusion of some of the constituent tumor cells. However, primary
lipase C activity, which results in intracellular calcium flux and tumor dormancy and metastatic dormancy seem to be different
downstream calcium-dependent cytoskeletal and/or transcrip- processes.145 In some cases, cells in the TME produce cytokines,
tional responses.129,130 such as Wnt proteins, secreted inhibitors of bone morphogenetic
Aberrant Wnt signaling is found in many cancers, such as protein (BMP), and Delta, which activate the signaling pathway to
invasive ductal breast carcinomas,131 colorectal cancer,132 papil- maintain the self-renewal ability of CSCs.146 Activation of Wnt
lary thyroid cancer,133 esophageal cancer,134 and colorectal induce the transformation of dormant CSCs into active CSCs to
cancer.135 The activation of Wnt signaling is different in different promote cell cycle progression through β-catenin, increasing the
tumors. Some Wnt activation is caused by mutations in Wnt expression of downstream cyclin D1 and MYC, and MYC also
components, such as Axin mutation in gastrointestinal cancers,136 promotes the expression of the polycomb repressor complex 1
APC mutation in colorectal cancer,137 and β-catenin mutation in component Bmi-1 and induces the combination E2F with cyclin
gastric cancer and liver cancer.138,139 GSK3 genes are critical for E.147 The extracellular matrix (ECM) protein tenascin C often exists
β-catenin regulation; therefore, many researchers expect the in the gap of stem cells, which supports the cell cycle in breast
occurrence of GSK3 mutations, but GSK3 mutations are not cancer cells by increasing Wnt signals.148 In addition, aberrant Wnt
correlated with cancer occurrence. In addition, some genes signaling has also been observed in the self-renewal of CSCs (Fig. 1).
(pyruvate kinase isozyme M2 (PKM2) in breast cancer140 and Many reports have proven that numerous proto-oncogenes
telomerase reverse transcriptase (TERT) in prostate cancer141) and stimulate this process through the Wnt signaling pathway.135
microRNAs (miR-164a in colorectal cancer142 and miR-582-3p in PKM2 catalyzes the last step of glycolysis and plays an essential
non-small-cell lung cancer143) inhibit the activity of APC, Axin, and role in the proliferation of breast CSCs by associating with
GSK3β to promote the accumulation of β-catenin in the increased β-catenin levels at regions “−410 to 180 and −2250 to
cytoplasm. 2000”.140,145,149 Enhancer of zeste homolog 2 (EZH2), a key
Stem cell signaling pathways and transcriptional circuits are component of the polycomb PRC2 complex, promotes self-
related to the alteration or reactivation of signaling pathways.144 renewal of CSCs by activating β-catenin.150 Moreover, TERT, an
Tumor dormancy is a lag phenomenon in tumor growth. RNA-dependent DNA polymerase, acts as a cofactor and forms a
Dormancy may occur during primary tumor formation or in the complex with β-catenin to activate Wnt downstream targets in

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
7
prostate CSCs.141 Capillary morphogenesis gene 2 increases the effectors, and Notch regulatory molecules. In 1917, studies
expression of nuclear β-catenin to regulate the self-renewal and discovered the Notch gene in a mutant Drosophila. Mammals
tumorigenicity of gastric CSCs,151 and SMYD3, which is located have four Notch receptors (Notch1–4) and five Notch ligands
downstream of the Wnt pathway, has a similar effect.152 In (Delta-like 1, 3, and 4, Jagged 1, and Jagged 2).178 Notch and DSL
addition, long noncoding RNAs and microRNAs also promote self- ligands are transmembrane proteins that mediate communication
renewal of CSCs through the Wnt signaling pathway. LncTCF7 between neighboring cells. Under physiological conditions, the
recruits the SWI/SNF complex to regulate the expression of the ligand binds to a Notch receptor that is expressed on neighboring
TCF7 promoter in liver CSCs.153 Lnc-β-Catm associates with the cells in a juxtacrine manner, thereby triggering proteolytic
methyltransferase EZH2 to suppress the ubiquitination of cleavage of the intracellular domain (ICD) of Notch and its
β-catenin and promote its stability,154 and LncTIC1 interacts with translocation into the nucleus to bind to the transcription factor
β-catenin and maintains its stability, activating Wnt/β-catenin CSL, forming the NICD/CSL transcriptional activation complex,
signaling.155 MicroRNA-1246, miR-19, and miR-92a suppress the which activates target genes of the bHLH transcription inhibitor
expression of AXIN and GSK3β in CSCs.156 MicroRNA-544a family, such as HES, HEY, and HERP.179,180
downregulates GSK3β in lung CSCs.157 MicroRNA-483-5p upregu- The Notch pathway regulates cancer cells in many tumors, such
lates the expression of β-catenin in gastric CSCs.158 In addition, as glioblastoma, leukemia, and those of the breast, pancreas,
there are still many genes, microRNAs, and noncoding RNAs in colon, and lung, among others.181 Different tumors and tumor
CSCs’ self-renewal through the Wnt signaling pathway. subtypes express different Notch ligands and receptors. Therefore,
Wnt signaling also plays an important role in the dedifferentia- Notch is known to function as both an oncogene and a
tion of CSCs. HOXA5, which is a member of the HOX family, suppressive gene. As an oncogene, Notch is overexpressed in
induces the differentiation of colorectal CSCs. However, Wnt gastric cancer,182 breast cancer,183 colon cancer,184 and pancreatic
indirectly suppresses indirectly via MYC, which is an important cancer. In contrast, Notch expression is downregulated in prostate
direct target of β-catenin/TCF in the intestine.159 PMP22, an cancer,185 skin cancer,186 non-small-cell lung cancer,187 liver
integral membrane glycoprotein in myelin in the peripheral cancer,188 and some breast cancers.189 Whether Notch acts as
nervous system, induces the differentiation of gastric CSCs, but its an oncogene or a tumor suppressor gene is determined by the
mRNA level declines with activation of the Wnt/β-catenin path- microenvironment.190 Moreover, post-translational modifications
way.160 Moreover, TRAP1, a component of the HSP90 (heat-shock of Notch receptors change their affinity for ligands and their
protein 90) chaperone family, inhibits the differentiation of intracellular half-lives.191
colorectal carcinoma stem cells by modulating β-catenin ubiqui- Many studies on the Notch pathway in CSCs have shown that
tination and phosphorylation.161 Lgr5, a member of the G protein- activation of Notch promotes cell survival, self-renewal, and
coupled receptor (GPCR) family of proteins, is located downstream metastasis and inhibits apoptosis. Aberrant Notch signaling
of the Wnt signaling pathway and restrains the differentiation of (Notch1 and Notch4) promotes self-renewal and metastasis of
esophageal SCC stem cells.162 breast and HCC stem cells.192,193 However, microRNA-34a down-
Wnt signaling also plays an important role in regulating CSC regulates Notch1.194 Similarly, abundant Delta-like ligand 4 (DLL4)
apoptosis. Dickkopf-related protein 2 induces G0/G1 arrest and also promotes tumor angiogenesis and metastasis in gastric
cell apoptosis by suppressing β-catenin activity in breast CSCs.163 CSCs.195 Delta-like 1 activation of Notch1 signaling requires the
DACT1, a homolog of Dapper that is located at chromosomal assistance of the actin-related protein 2/3 complex to maintain the
region 14q23.1, promotes apoptosis in breast CSCs by antagoniz- stem cell phenotype of glioma-initiating cells.196 Additionally,
ing the Wnt/β-catenin signaling pathway.164 Cadherin-11, a some intracellular genes also regulate the Notch signaling
proapoptotic tumor suppressor, reduces the level of active pathway. For example, MAP17 (DD96, PDZKIP1), a nonglycosylated
phospho-β-catenin (ser552) to induce apoptosis in colorectal membrane-associated protein, is located on the plasma mem-
CSCs.165 Epigallocatechin-3-gallate increases apoptosis by degrad- brane and the Golgi apparatus. MAP17 interacts with NUMB
ing β-catenin in lung CSCs.166 The small-molecule inhibitor through the PDZ-binding domain to activate the Notch pathway
CWP232228 antagonizes the binding of β-catenin to TCF in the in cervical CSCs.197 Inducible nitric oxide synthase promotes the
nucleus to induce apoptosis in liver CSCs.167 In addition, self-renewal capacity of CD24+CD133+ liver CSCs through TACE/
temozolomide combined with miR-125b significantly induces ADAM17 activation to regulate Notch1 signaling.198 Moreover,
apoptosis by targeting the Wnt/β-catenin signaling pathway in tumor necrosis factor-α (TNFα) enhances the CSC-like phenotype
glioma stem cells.168 by activating Notch1 signaling in oral SCC cells.199 Overexpression
Wnt/β-catenin signaling has been implicated in CSC-mediated of PER3 decreases the expression of Notch1 and Jagged 1 in
metastasis.169 In the cytomembrane, Frizzled8 promotes bone colorectal CSCs.200 In addition, KLF4 and BMP4 also increase
metastasis in prostate CSCs.170 The leucine-rich repeat containing Notch1 and Jagged 1 in breast CSCs to regulate cell migration and
GPCR4 (LGR4, or GPR48), together with its family members LGR5/ invasion.201,202 BRCA1 is a key regulator of breast cancer cell
6, binds to R-spondins 1–4 and leads to Wnt3A potentiation, differentiation; however, it is localized to a conserved intronic
activating Wnt signaling in breast CSCs.171,172 Increased levels of enhancer region within the Notch ligand Jagged 1 gene to
CD44v6 mRNA in human pancreatic CSCs, lung CSCs, and colon maintain the stemness of breast CSCs.203 Similarly, increased Gli3
CSCs promote migration and metastasis through the activation of also promotes cell proliferation and invasion in oral SCC by
β-catenin.173–175 In the cytoplasm, TAZ/YAP interacts directly with increasing Notch2.204 Hypoxia/hypoxia-inducible factor (HIF)-
β-catenin and restricts β-catenin degradation,176 but TIAM1 induced migration and invasion is a well-known phenomenon
antagonizes TAZ/YAP accumulation and translocation from the that has been reported in numerous CSCs.205 Notch1 can induce
cytoplasm to the nucleus.177 Moreover, CDH11 inhibits the the migration and invasion of ovarian CSCs in the absence of
migration and invasion of colorectal CSCs by inhibiting Wnt/ hypoxia.206 Hypoxia-induced Jagged 2 activation enhances cell
β-catenin and AKT/RhoA signaling.165 Wnt signaling decreases the invasion of breast CSCs207 and lung CSCs.208 Moreover, HIF-1α/2α
expression of HOXA5 to promote CSC metastasis.159 These data regulates self-renewal and maintenance of glioblastoma stem
suggest that amplified Wnt signaling is important for self-renewal, cells.209 In addition, increased miR-200b-3p decreases Notch
dedifferentiation, apoptosis inhibition, and metastasis of CSCs. signaling to promote pancreatic CSCs to become asymmetric.210
MiR-26a directly targets Jagged 1 to inhibit osteosarcoma CSC
Notch signaling pathway in CSCs. The Notch signaling pathway proliferation.211 These studies indicate that Notch plays an
consists of the Notch receptor, Notch ligand (DSL protein), CSL important role in regulating the self-renewal, growth, and
(CBF-1, suppressor of hairless, Lag), DNA-binding protein, other metastasis of CSCs.

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
8
Hh signaling pathway in CSCs. The Hh signaling pathway consists the transcription factor GLI and inducing it to enter the cell
of ligands and receptors. The Hh signaling network is very nucleus, where GLI regulates cell growth, proliferation, and
complex, including extracellular Hh ligands, the transmembrane differentiation.221
protein receptor PTCH, the transmembrane protein SMO, inter- Studies have confirmed that abnormal activation of the Hh
mediate transduction molecules, and the downstream molecule signaling pathway can be found in human cancers,222 such as
GLI.212 The components of the Hh signaling pathway play different breast cancer,223 lung cancer,224 bladder cancer,225 pancreatic
roles. The membrane protein SMO plays a positive regulatory role, cancer,226 chondrosarcoma,227 rhabdomyosarcoma,228 neuroblas-
while the transmembrane protein PTCH plays a negative toma,229 medulloblastoma,230 and gastric cancer.231 However,
regulatory role. PTCH has two subtypes, PTCH1 and PTCH2,213 activation of Hh signaling is different in different tumors. Gorlin
and there is 73% homology between the two subtypes. GLI, an syndrome (basal cell nevus syndrome), an autosomal dominant
effector protein, has three subtypes, Gli1, Gli2, and Gli3, in condition, is associated with germline loss of the PTCH1 gene. This
vertebrates,214 and these effector proteins have different func- condition is very common in basal cell carcinoma, rhabdomyo-
tions. Gli1 strongly activates transcription, while Gli3 inhibits sarcoma, and medulloblastoma.232,233 Other Hh pathway compo-
transcription.215 Gli2 has dual functions of activating and nents are also mutated in human cancers, such as Gli1 and Gli3
inhibiting transcription but mainly functions as a transcriptional mutations in pancreatic adenocarcinoma, Gli1 gene amplification
activator.216,217 Numerous studies have confirmed that Hh in glioblastoma, and SUFU (suppressor of fused) mutations in
signaling is involved in embryonic development and the medulloblastoma.234,235 In addition, other genes also regulate the
formation of the nervous system, skeleton, limbs, lung, heart, Hh signaling pathway. Speckle-type POZ protein, an E3 ubiquitin
and gut.218 As an extracellular signaling pathway, in the presence ligase adaptor, inhibits Hh signaling by accelerating Gli2
of ligand signals, Hh ligands bind to PTCH receptors on target cell degradation in gastric cancer.236
membranes and initiate a series of intracellular signal transduction Hh signaling plays distinct functions in different types of
processes.219 When there is no ligand signal, the transmembrane cancer.237 During tumor development, Hh signaling has three
receptor PTCH on the target cell membrane binds to SMO and major roles: driving tumor development, promoting tumor
inhibits SMO activity, which prevents signaling.220 When the Hh growth, and regulating residual cancer cells after therapy. Based
ligand is present, it binds to PTCH, which changes the spatial on these functions, the aberrant Hh pathway plays a causal role in
conformation of PTCH, removing the inhibition of SMO activating CSCs238,239 (Fig. 2). The expression level of Hh signaling

Fig. 2 Hedgehog signaling pathway in cancer stem cells. The Hedgehog pathway plays a key role in stem maintenance, self-renewal, and
regeneration of CSCs. The secreted Hh protein acts in a concentration- and time-dependent manner to initiate a series of cell responses, such as
cell survival, proliferation, and differentiation. After receiving the Shh signal, the transmembrane protein receptor PTCH relieves the inhibition of
the transmembrane protein SMO, which induces Gli1/2 to detach from SUFU and enter the nucleus to regulate downstream gene transcription.
During activation of the Hh pathway, some proteins (IL-6, IL-27, Fbxl17 (F-box and leucine-rich repeat protein 17), PPKCI, RARα2, RUXN3, SCUBE2,
HDAC6 (histone deacetylase 6), USP48, CK2α, WIP1, GALNT1, VASH2 (Vasohibin 2), BCL6, FOXC1 (forkhead box C1), and p65), microRNAs (miR-324-
5p, miR-122, and miR-326), and the long noncoding RNA HDAC2 are involved in the Hedgehog pathway to affect CSC growth

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
9
components is relatively high in CSCs. For example, Hh signaling NF-κB signaling pathway in CSCs. Nuclear factor-κB (NF-κB), a
promotes the maintenance, proliferation, self-renewal, and rapidly inducible transcription factor,274 consists of five different
tumorigenicity of lung adenocarcinoma stem cells.240 In CD133+ proteins (p65, RelB, c-Rel, NF-κB1, and NF-κB2). The main
glioma stem cells, SMO, GLI, and PTCH promote cell proliferation, physiological function of NF-κB is the p50-p65 dimer.275–277 The
self-renewal, migration, and invasion. The expression of Gli1, primary mode of NF-κB regulation occurs at the level of
PTCH1, and PTCH2 is regulated by histone deacetylase 6.241 USP48 subcellular localization. In the activation stage, transcription factor
activates Gli-dependent transcription by stabilizing the Gli1 complexes must translocate from the cytoplasm to the nucleus.278
protein in glioma stem cells.242 The protein kinase CK2α enhances The activity of the complexes is regulated by two major pathways
Gli1 expression and its transcriptional activity in lung CSCs.243 (canonical NF-κB signaling and noncanonical NF-κB signaling). In
WIP1 (PPM1D), a nuclear Ser/Thr phosphatase, also enhances the the canonical NF-κB activation pathway, activation occurs through
function of Gli1 by increasing its transcriptional activity, protein the binding of ligands, such as bacterial cell components, IL-1β,
stability, and nuclear localization in breast CSCs and medullo- TNF-α, or lipopolysaccharides, to their respective receptors, such
blastomas.244,245 F-box and leucine-rich repeat protein 17 as Toll-like receptors, TNF receptor (TNFR), IL-1 receptor (IL-1R),
mediates the release of Gli1 from SUFU for proper Hh signal and antigen receptors.279 Stimulation of these receptors leads to
transduction in medulloblastoma stem cells.246 Moreover, retinoic the phosphorylation and activation of IκB kinase (IKK) proteins,
acid receptor α2 (RARα2) upregulates the expression of SMO and subsequently initiating the phosphorylation of IκB proteins.276 The
Gli1 in CD138+ multiple myeloma stem cells.247 PRKCI, which is alternative pathway of NF-κB activation is termed the noncano-
regulated by miR-219 in tongue SCC,248 has a similar function as nical pathway. The noncanonical pathway receptor originates
RARα2 in maintaining a stem-like phenotype in lung SCC cells.249 from different classes, such as CD40, receptor activator for NF-κB,
Interleukin-27 (IL-27) and IL-6 activate Hh signaling in CD133+ B cell activation factor, TNFR2 and Fn14, and lymphotoxin
non-small-cell lung CSCs.250 During self-renewal and maintenance β-receptor.280 This pathway leads to activation of NF-κB by
of stemness of BCMab1+CD44+ bladder CSCs, glycotransferase inducing the kinase (NIK), which then phosphorylates and
GALNT1-mediated glycosylation significantly activates Sonic Hh predominantly activates IKK1. The activity of the latter enzyme
signaling by upregulating Gli1.251 induces the phosphorylation of p100 to generate p52.281
Furthermore, p63, a master regulator of normal epithelial stem The NF-κB pathway plays an important role in regulating
cell maintenance, regulates the expression of Shh, Gli2, and immune and inflammatory responses. In addition, the NF-κB
PTCH1 by directly binding to their gene regulatory regions, which pathway is involved in cellular survival, proliferation, and
eventually contributes to the activation of Hh signaling in differentiation.276 The process of tumor development and
mammary CSCs.252 The N-terminal domain of forkhead box C1 progression produces cytokines, growth, and angiogenic factors
binds directly to an internal region (amino acids (aa) 898–1168) of and proteases to activate NF‐κB signaling.282 Inflammation has
Gli2 to enhance transcriptional activation of Gli2 and determines been recognized as a hallmark of cancer.283 Overactivation of NF-
the stem cell phenotype in breast CSCs.253 Through recruitment κB signaling has been reported in gastrointestinal, genitourinary,
of the deubiquitinating enzyme ATXN3, tetraspanin-8 interacts gynecological, and head and neck cancers, breast tumors, multiple
with PTCH1 and inhibits the degradation of the SHH/PTCH1 myeloma, and blood cancers.278,284–286 However, direct or altered
complex. In addition, long noncoding microRNAs also activate Hh molecular mutations in NF-κB have rarely been reported in human
signaling. For example, lncHDAC2 promotes the self-renewal of cancers.287 Based on recent studies, NF-κB regulates many genes
liver CSCs by recruiting the NuRD complex onto the promoter of and is implicated in cell survival, proliferation, metastasis, and
the PTCH1 gene to suppress its expression.254 In addition, the TME tumorigenesis of cancer.288 NF-κB activation also directly or
is crucial for the survival of CSCs. Consequently, breast CSCs indirectly enhances the expression of key angiogenesis factors
secrete Shh, which upregulates cancer-associated fibroblasts and adhesion molecules, such as IL-8, vascular endothelial growth
(CAFs). Subsequently, CAFs secrete factors that promote the factor (VEGF), and growth-regulated oncogene 1.289
expansion and self-renewal of breast CSCs.255 Hh signaling also The NF-κB pathway has an essential connection regulating
promotes self-renewal and metastasis of CSCs by upregulating inflammation, self-renewal, or maintenance and metastasis of
the expression of related downstream markers of CSCs, such as CSCs (Fig. 3). CD44+ cells promote self-renewal, metastasis, and
Bmi-1, Wnt2, ALDH1, CD44, CCND1, Twist1, C-MYC, Nanog, Oct4, maintenance of ovarian CSCs by increasing the expression of RelA,
PDGFRα (platelet-derived factor receptor-α), Snail, Jagged 1, and RelB, and IKKα and mediating nuclear activation of p50/RelA (p50/
C-MET.231,247,256–264 p65) dimer.290 High levels of NIK induce activation of the
Some proto-oncogenes and suppressor genes also directly or noncanonical NF-κB pathway to regulate the self-renewal and
indirectly regulate Hh signaling in the proliferation and migration metastasis of breast CSCs.291 Moreover, stromal cell-derived
of CSCs. The signal peptide CUB EGF-like domain-containing factor-1 (SDF-1) also has the same effect by regulating the
protein 2 (SCUBE2), a member of the SCUBE family of proteins, translocation of p65 from the cytoplasm to the nucleus.292 The
inhibits cell proliferation and migration in glioma stem cells by inflammatory mediator prostaglandin E2 (PGE2) contributes to
downregulating Hh signaling.265 BCL6, a transcriptional repressor tumor formation, maintenance, and metastasis by activating NF-
and lymphoma oncoprotein, directly represses the Sonic Hh κB via EP4-PI3K (phosphoinositide 3-kinase) and EP4-MAPK
effectors Gli1 and Gli2 in medulloblastoma stem cells.266 The pathways in colorectal CSCs.293 Chemokines, low-molecular-
transcription factor RUNX3 suppresses metastasis and the weight proinflammatory cytokines, are important mediators of
stemness of colorectal CSCs by promoting ubiquitination of Gli1 cell proliferation, metastasis, and apoptosis.294 C-C chemokine
at the intracellular level.267 Vasohibin 2 suppresses Smo, Gli1, and receptor 7 interacts with its ligand chemokine ligand 21 to inhibit
Gli2 expression in pancreatic CSCs.268 β-Catenin stably increases apoptosis and induce survival and migration in CD133+ pancreatic
its physical interaction with Gli1, resulting in Gli1 degradation in cancer stem-like cells by increasing the expression of extracellular
medulloblastoma stem cells.269 In addition, microRNAs also target signal-regulated kinase 1/2 (Erk1/2) and p65.295 Furthermore, B
Hh signaling components to regulate CSC proliferation. For cell-specific Moloney murine leukemia virus integration site 1
example, miR-324-5p significantly decreases SMO and Gli1 in (Bmi-1) also enhances the p65 protein in gastric CSCs.296
myeloma stem cells.270 Mir-326 directly downregulates SMO and MicroRNAs also play an important role in promoting the
Gli2 in medulloblastoma stem cells.271 MiR-326 downregulates proliferation of CSCs. Mir-221/222 promotes self-renewal, migra-
SMO in glioma stem cells.272 Mir-122 targets Shh and Gli1 in lung tion, and invasion in breast CSCs by inhibiting the expression of
CSCs.273 These data demonstrate that amplified Hh signaling is PTEN and then inducing the phosphorylation of AKT, resulting in
important for the self-renewal, growth, and metastasis of CSCs. elevated p65, p-p65, and COX2.297

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
10

Fig. 3 NF-κB signaling pathway in cancer stem cells NF-κB proteins are involved in the dimerization of transcription factors, regulate gene
expression, and affect various CSC biological processes, including inflammation, stress responses, growth, and development of CSCs. The
main physiological function of NF-κB is the p50-p65 dimer. The active p50-p65 dimer is further activated by post-translational modification
(phosphorylation, acetylation, or glycosylation) and transported into the nucleus, which induces the expression of target genes in
combination with other transcription factors. Some proteins (CD44, CD146, TNFRSF19, Bmi-1, FOXP3, and SDF-1) and microRNAs (miR-221 and
miR-222) directly regulate the NF-κB pathway. In addition, some proteins (PGE2, GIT-1 (G protein-coupled receptor kinase-interacting protein
1), C-C chemokine receptor 7 (CCR7), and TGF-β) and miR-491 indirectly affect the NF-κB pathway via the ERK and MAPK pathways in CSCs

In addition, other transcription factors also inhibit self-renewal three components: the tyrosine kinase-related receptor, the
and metastasis in CSCs by the NF-κB pathway. Increased tyrosine kinase JAK, and the transcription factor STAT.304 Many
expression of FOXP3 has been identified in different cancers.298 cytokines and growth factors transmit signals through the JAK-
FOXP3 interacts with NF-κB, inhibits the expression of COX2 STAT signaling pathway, including interleukin-2-7, granulocyte/
located downstream of NF-κB, and affects self-renewal and macrophage colony-stimulating factor, growth hormone, EGF,
metastasis in colorectal CSCs.299 Overexpression of miR-491 blocks PDGF, and interferon.305 These cytokines and growth factors have
the activation of NF-κB in liver CSCs by targeting G protein- corresponding receptors on the cell membrane. The common
coupled receptor kinase-interacting protein 1, which inhibits characteristic of these receptors is that the receptor itself does not
ERKs.300 Moreover, some drugs inhibit cell proliferation and have kinase activity, but there is a binding site for the tyrosine
metastasis of CSCs by the NF-κB pathway. Disulfiram, an anti- kinase JAK in the cells. After binding with ligands, tyrosine
alcoholism drug, inhibits tumor growth factor-β (TGF-β)-induced residues of various target proteins are phosphorylated through
metastasis via the ERK/NF-κB/Snail pathway in breast CSCs.301 JAK activation to achieve signal transduction from the extracellular
Sulforaphane preferentially inhibits self-renewal in triple-negative to intracellular space. The JAK protein family consists of four
breast CSCs by inhibiting NF-κB p65 subunit translocation and members: JAK1, JAK2, JAK3, and Tyk2.306 JAK proteins have seven
downregulating p52 and its transcriptional activity.302 Curcumin JAK homology (JH) domains in their structures. The JH1 domain is
regulates the proliferation, metastasis, and apoptosis of HCC stem the kinase domain, the JH2 domain is the "pseudo" kinase domain,
cells by inhibiting the NF-κB pathway.303 These data demonstrate and JH6 and JH7 are the receptor binding domains.307 STAT is
that amplified NF-κB signaling is important for regulating called "signal transducer and activator of transcription". As the
apoptosis, proliferation, and metastasis of CSCs. name implies, STAT plays a key role in signal transduction and
transcriptional activation. At present, seven members of the STAT
JAK-STAT signaling pathway. The Janus kinase/signal transducers family (STAT1, STAT2, STAT3, STAT4, STAT5a, STAT5b, STAT6) have
and activators of transcription (JAK-STAT) signaling pathway is a been identified. The structure of STAT protein can be divided into
signal transduction pathway that is stimulated by cytokines. This the following functional regions: N-terminal conserved sequence,
pathway is involved in many important biological processes, such DNA-binding region, SH3 domain, SH2 domain, and C-terminal
as cell proliferation, differentiation, apoptosis, and immune transcriptional activation region.308 Generally, many cytokines and
regulation. Compared with the complexity of other signaling growth factors integrate with tyrosine kinase-related receptors.
pathways, this signaling pathway is relatively simple. There are After receiving the signal from the upstream receptor molecule,

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
11
JAK is quickly recruited to and activates the receptor, resulting in type II receptor, which recruits a type I receptor and phosphor-
JAK activation to catalyze tyrosine phosphorylation of the ylates it. This type I receptor phosphorylates receptor-regulated
receptor. The phosphorylated tyrosine on the receptor molecule, Smads (R-Smads), which bind to common pathway Smad (co-
which is a signaling molecule, can bind with the SH2 site of Smad). The R-Smad/co-Smad complex acts as a transcription
STAT.309 When STAT binds to the receptor, tyrosine phosphoryla- factor and accumulates in the nucleus to regulate the expression
tion of STAT also occurs, which forms a dimer and enters the of target genes. TGF-β superfamily ligands include BMPs, growth
nucleus.310 As an active transcription factor, the STAT dimer and differentiation factors (GDFs), anti-Mullerian hormone (AMH),
directly affects the expression of related genes and then changes activin Nodal, and TGF-β.339 These ligands can be divided into two
the proliferation or differentiation of target cells.311 groups, TGF-β/activin and BMP/GDF. The TGF-β/activin group
Constitutive activation of JAKs and STATs was first recognized includes TGF-β, activin, and Nodal, and the BMP/GDF group
as being associated with malignancy in the 1990s.312 Based on includes BMP, GDF, and AMH ligands.340 Based on Smad structure
current studies, JAK2 mutation and abnormal activation of STAT3 and functions, Smad proteins can be divided into three
are prone to occur in many tumors.313 Mutations in JAK2 have subfamilies: receptor-activated or pathway-restricted Smad (R-
been reported in the majority of patients with myeloproliferative Smads), Co-Smad, and inhibitory Smad (I-Smads), which includes
neoplasms,314 such as polycythemia vera, myelofibrosis, and at least nine Smad proteins.341,342 R-Smads are activated by type I
thrombocythemia.315,316 These disorders are caused by the receptors and form transient complexes with these receptors.
overexpansion of hematopoietic precursors, which are often There are two types of Smad complexes: AR-Smads are activated
clonal and can result in leukemia.314 Several lines of evidence by activin TGF-β, including Smad2 and Smad3, and BR-Smads are
show that constitutive activation of JAK2 and STAT3 in the activated by BMP, including Smad1, Smad5, Smad8, and Smad9.
absence of any stimulating ligand occurs in polycythemia Co-Smad, including Smad4, is a common medium in various TGF-β
vera.317,318 Moreover, studies have also found aberrant activation signal transduction processes. I-Smads, including Smad6 and
of STATs in human cancers, such as head and neck cancer,319 Smad7, bind to activated type I receptors and inhibit or regulate
endometrial cancer,320 breast cancer, diffuse large B cell signal transduction of the TGF-β family.343
lymphoma,321 HCC,322 colorectal cancer, glioma,323 and colon Many studies have shown that activation of TGF/Smad signaling
cancer.324 Furthermore, aberrant STAT5 signaling has been found also occurs in human cancers. Dkk-3, a secreted protein, inhibits
in the pathogenesis of hematologic and solid organ malignan- TGF-β-induced expression of matrix metallopeptidase 9 (MMP9)
cies.325,326 and MMP13 to prevent migration and invasion of prostate
The JAK/STAT pathway is evolutionarily conserved. This path- cancer.344 Cancer upregulated gene 2 promotes cellular transfor-
way promotes the survival, self-renewal, hematopoiesis, and mation and stemness, which is mediated by nuclear NPM1 protein
neurogenesis of ESCs.327 This pathway is also activated in CSCs. and TGF-β signaling in lung cancer.345 TGF/Smad also plays an
The persistent activation of STAT3 significantly promotes cell important role in the cell proliferation of CSCs. Cyclin D1 interacts
survival and the maintenance of stemness in breast CSCs.328 IL-10 with and activates Smad2/3 and Smad4, promoting cyclin D1-
induces cell self-renewal, migration, and invasion in non-small-cell Smad2/3-Smad4 signaling to regulate self-renewal of liver CSCs.346
lung CSCs.329 IL-6 activates the JAK1/STAT3 pathway in ALDHhigh CD51 binds to TGF-β receptors to upregulate TGF-β/Smad
CD126+ endometrial CSCs.320 Furthermore, IL-6 also induces the signaling in colorectal CSCs.341 Upregulation of TGF-β1 induces
conversion of nonstem cancer cells into cancer stem-like cells in the expression of smad4, p-Smad2/3, and CD133 in liver CSCs.347
breast cancer by the activating downstream Oct4 gene.330 Oct4 TGF-β1 also upregulates the expression of PFKFB3 through
also activates the JAK1/STAT6 pathway in ovarian CSCs.331 In activation of the p38 MAPK and PI3K/Akt signaling pathways to
CD44+CD24− breast and colorectal CSCs, erythropoietin, and IL-6 regulate glycolysis in glioma stem cells.348 Furthermore, silencing
activate the JAK2/STAT3 pathway.332–334 Retinol-binding protein 4 ShcA expression also induces activation of STAT4 in breast
activates JAK2/STAT3 signaling by its STRA6 receptor in colon CSCs.349 Moreover, miR-148a inhibits the TGF-β/Smad2 signaling
CSCs.319 HIF-1α enhances the self-renewal of glioma stem-like cells pathway in HCC stem cells.350 Smad7, a newly discovered target
by the JAK1/STAT3 pathway.335 AJUBA is a scaffold protein that gene of miR-106b, is an inhibitor of TGF-β/Smad signaling, which
participates in the regulation of cell adhesion, differentiation, inhibits sphere formation of gastric cancer stem-like cells.351
proliferation, and migration and promotes the survival and Although there are few studies on the TGF/Smad signaling
proliferation of colorectal CSCs via the JAK1/STAT1 pathway.336 pathway in CSCs, this pathway still plays a very important role.
Moreover, microRNAs are also involved in activating JAK/STAT
signaling by inhibiting negative regulatory factors of JAK2/STAT3. PI3K/AKT/mTOR signaling pathway in CSCs. Phosphatidylinositol-
For example, miR-500a-3p targets multiple negative regulators of 3-kinase (PI3K) is an intracellular phosphatidylinositol kinase.352 It
the JAK2/STAT3 signaling pathway, such as SOCS2, SOCS4, and consists of the regulatory subunit p85 and catalytic subunit p110,
PTPN, in HCC stem cells, leading to constitutive activation of which have serine/threonine (Ser/Thr) kinase and phosphatidyli-
STAT3 signaling.322 MiR-30 targets SOCS3 in glioma stem cells.337 nositol kinase activities.353 AKT is a serine/threonine kinase that is
Mir-93 downregulates the expression of JAK1 and STAT3 to induce expressed as three isoforms: AKT1, AKT2, and AKT3.354 AKT
the differentiation of breast CSCs. Mir-218 negatively regulates the proteins are crucial effectors of PI3K and are directly activated in
IL-6 receptor and JAK3 gene expression in lung CSCs.338 In response to PI3K. One of the key downstream target genes of AKT
addition, some endogenous or exogenous genes inhibit JAK/STAT is the mammalian target of rapamycin (mTOR) complex, which is a
signaling in CSCs. Von Hippel–Lindau suppresses the tumorigeni- conserved serine/threonine kinase. It forms two distinct multi-
city and self-renewal ability of glioma stem cells by inhibiting protein complexes: mTORC1 and mTORC2.355 mTORC1 consists of
JAK2/STAT3.323 Although there are few studies on JAK in CSCs, mTOR, raptor, mLST8, and two negative regulators, PRAS40 and
there is a role for JAK/STAT signaling in the survival, self-renewal, DEPTOR.356,357 mTORC2 phosphorylates AKT at serine residue 473,
and metastasis of CSCs. which leads to full AKT activation.358
Studies show that mutations in PTEN lead to the inhibition of
TGF/SMAD signaling pathway in CSCs. The TGF-β signaling PI3K/mTOR signaling in glioblastoma multiforme. However,
pathway is involved in many cellular processes associated with deletion of PTEN in neural stem cells leads to a neoplastic
organism and embryo development, including cell proliferation, phenotype that includes cell growth promotion, resistance to cell
differentiation, apoptosis, and homeostasis. Although the TGF-β apoptosis, and increased migratory and invasive properties
signaling pathway regulates a wide range of cellular processes, its in vivo.359 Inactivation of PTEN and activation of protein kinase
structure is relatively simple. TGF-β superfamily ligands bind to a B have been found in other solid tumors, such as

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
12
myeloproliferative neoplasia and leukemia.360 Therefore, the PI3K/ canonical response element in the miR-15a gene in breast CSCs to
mTOR signaling pathway is vital for cell proliferation and survival. reduce the CD49high/CD24+ mesenchymal stem cell (MSC)
Abnormal activation of PI3K/mTOR signaling is found in some population and inhibit angiogenesis.386 PPARγ activation also
cancers, such as non-small-cell lung cancer,361 breast cancer,362 prevents cell spheroid formation and stem cell-like properties in
prostate cancer,363 Burkitt lymphoma,364 esophageal adenocarci- bladder CSCs and induces adipocyte differentiation and β-catenin
noma,365 and colorectal cancer.366 degradation in adipose tissues.387 Furthermore, expression of
Although PI3K/AKT/mTOR has been extensively studied in PPARγ restrains YAP transcriptional activity to induce differentia-
cancers, there are few studies in CSCs.358 PI3K/Akt/mTOR signaling tion in osteosarcoma stem cells388 and melanoma cells.389 The
is involved in ovarian cancer cell proliferation and the PPARγ/NF-κB pathway promotes M2 polarization of macrophages
epithelial–mesenchymal transition.367 This signaling activation to prevent cell death in ovarian CSCs4.390 PPARγ activation
also enhances the migration and invasion of prostate and promotes expression of its target gene PTEN to inhibit PI3K/Akt/
pancreatic CSCs.368,369 Downregulation of PTEN induces PI3K mTOR signaling, which stunts self-renewal, tumorigenicity, and
activation to promote survival, maintenance of stemness, and metastasis in cervical CSCs, glioblastoma stem cells, and liver
tumorigenicity of CD133+/CD44+ prostate cancer stem-like cell CSCs.391,392 However, combined expression of Dnmt3a and
populations.370 PI3K activation promotes cell proliferation, migra- Dnmt3b inhibits PPARγ expression by direct methylation of its
tion, and invasion in ALDH+CD44high head and neck squamous promoter in squamous carcinomas.393 PPARs are also closely
CSCs.371 Activation of mTOR promotes the survival and prolifera- related to the metabolism of CSCs. PPARα and PPARβ/δ regulate
tion of breast CSCs and nasopharyngeal carcinoma stem metabolic reprogramming in glioblastoma stem cells, lung CSCs,
cells.328,372 mTORC1 activation also increases aldehyde dehydro- and mouse mammary gland cancer.394 The transcription coacti-
genase 1 (ALDH1) activity in colorectal CSCs.373 Activation of vator peroxisome proliferator-activated receptor gamma coacti-
mTORC2 upregulates the expression of the hepatic CSC marker vator 1α (PPARGC1A, also known as PGC-1α) promotes CSC
EpCAM (epithelial cellular adhesion molecule) and tumorigenicity proliferation and invasion by enhancing oxidative phosphoryla-
in hepatocellular CSCs.374 Nucleotide-binding domain and leucine- tion, mitochondrial biogenesis, and the oxygen consumption rate
rich repeats (NLRs) belong to a large family of cytoplasmic sensors. of breast CSCs.395 In addition, the AMPK signaling pathway
NLRC3 (also known as CLR16.2 or NOD3) is associated with PI3Ks (adenosine 5′-monophosphate (AMP)-activated protein kinase) is
and blocks activation of PI3K-dependent kinase AKT in colorectal an AMP-dependent protein kinase that is a key molecule in the
CSCs.375 regulation of bioenergy metabolism and is the core of the study of
In addition, some studies have shown that the mTOR signaling diabetes and other metabolic-related diseases. AMPK is expressed
pathway is closely related to the metabolism of CSCs. For example, in various CSCs related to metabolism. Some studies have shown
low folate (LF) stress reprograms metabolic signals through the that AMPK is necessary for prostate CSCs to maintain glucose
activated mTOR signaling pathway, promoting the metastasis and balance.396 Metformin, an antidiabetic drug that fights cancer,
tumorigenicity of lung cancer stem-like cells.376 However, matcha targets AMPK signaling to inhibit cell proliferation and metabolism
green tea (MGT), an inhibitor of mTOR, inhibits the proliferation of in colorectal CSCs397 and HCC stem cells.398 Therefore, metformin
breast CSCs by targeting mitochondrial metabolism, glycolysis, may be a potential therapeutic regent by regulating the energy
and multiple cell signaling pathways.377 A link between the PI3K/ metabolism of CSCs. These studies suggest that PPARs play an
Akt/mTOR pathway and CSCs is clearly evident. important role in the growth of CSCs.

PPAR signaling pathways in CSCs. Peroxisome proliferator- Interactions between signaling pathways in CSCs. As mentioned
activated receptors (PPARs) are ligand-activated nuclear transcrip- previously, these complex signal transduction pathways are not
tion factors that were first cloned from mouse liver by Isseman linear. In some cases, crosstalk between and among various
and Green.378 PPARs are also members of the ligand-activated pathways occurs to regulate CSCs.399 Wnt/β-catenin and NF-κB
transcription factor superfamily of nuclear hormone receptors that signaling work together to promote cell survival and proliferation of
are associated with retinoic acid, steroids and thyroid hormone CSCs. TNFRSF19, a member of the TNF receptor superfamily, is
receptors. PPARs act as fat sensors to regulate the transcription of regulated in a β-catenin-dependent manner, but its receptor
lipid metabolic enzymes.379 At present, three subtypes, PPARα, molecules activate NF-κB signaling to regulate the development of
PPARβ, and PPARγ (encoded by the PPARA, PPARD, and PPARG colorectal cancer.400 Knockdown of CD146 results in inhibition of NF-
genes, respectively), have been found.380 PPARα is highly κB/p65-initiated GSK3β expression, which promotes nuclear translo-
expressed in hepatocytes, cardiac myocytes, intestinal cells, and cation and activation of β-catenin.401 In addition, there is negative
renal proximal convoluted tubule cells. PPARγ is abundantly regulation between Wnt/β-catenin and NF-κB signaling. Studies
expressed in adipose tissue, vascular parietal cells (such as have revealed a negative effect of β-catenin on NF-κB activity in
monocytes/macrophages, ECs, and smooth muscle cells), and liver, breast, and colon cancer cells.402,403 Leucine zipper tumor
myocardial cells.381 PPARβ is expressed in almost all tissues of the suppressor 2 (LZTS2) is a putative tumor suppressor, and NF-κB
body, and its expression level is higher than that of PPARα or activation inhibits β-catenin/TCF activity through upregulation of
PPARγ.382 In recent years, studies have found that PPARs are LZTS2 in liver, colon, and breast cancer cells.404–406 Wnt/β-catenin
closely related to energy (lipid and sugar) metabolism, cell and Hh signaling have important functions in embryogenesis, stem
differentiation, proliferation, apoptosis, and inflammatory reac- cell maintenance, and tumorigenesis. Wnt/β-catenin signaling
tions.383 PPARs can exert positive or negative effects to regulate induces the expression of CRD-BP, an RNA-binding protein, which
target gene expression by binding to a specific peroxisome results in the binding and stabilization of Gli1 mRNA, leading to an
located at each gene regulatory site and a proliferative response increase in Gli1 expression and transcriptional activity, which
element.378 Their natural ligands are unsaturated fatty acids, promotes the survival and proliferation of colorectal CSCs.407
eicosane acids, oxidized low-density lipoprotein, very low-density However, a report showed that noncanonical Hh signaling is a
lipoprotein, and linoleic acid derivatives.384 positive regulator of Wnt signaling in colon CSCs.408
To date, there have been many reports about the role of PPARs In addition, crosstalk between pathways promotes cell growth
in cancer cells, including prostate cancer, breast cancer, glioblas- and metastasis through maintenance of the CSC population.
toma, neuroblastoma, pancreatic cancer, hepatic cancer, leukemia, Downregulation of Notch1 and IKKα enhances NF-κB activation to
and bladder cancer and thyroid tumors.385 However, the function promote the CD133+ cell population in melanoma CSCs.409 IL-6/JAK/
of PPARs in CSCs is not well understood, except for some reports STAT3 and TGF-β/Smad signaling induce the proliferation and
on PPARγ. As a tumor suppressor, PPARγ binds and activates a metastasis of lung CSCs.410 IL-17E binding to IL-17RB activates the

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
13

Fig. 4 The microenvironment of cancer stem cells. Proliferation, self-renewal, differentiation, metastasis, and tumorigenesis of CSCs in the CSC
microenvironment. The CSC microenvironment is mainly composed of vascular niches, hypoxia, tumor-associated macrophages, cancer-
associated fibroblasts, cancer-associated mesenchymal stem cells, and extracellular matrix. These cells in response to hypoxic stress and
matrix induce growth factors and cytokines (such as IL-6 and VEGF) to regulate the growth of CSCs via Wnt, Notch, and other signaling
pathways

NF-κB and JAK/STAT3 pathways to promote proliferation and sustain that is directly involved in the microcirculation of tumors.421,422 ECs
self-renewal of CSCs in HCC.411 TGF-β1 silencing decreases the also promote CSC-like transformation and cell growth through Shh
expression of Smad2/3, β-catenin, and cleaved-Notch1 to inhibit the activation of Hh signaling.423 Moreover, secreted microvesicles of
activation of Wnt and Notch signaling in liver CSCs.346 Activation of CSCs promote the proliferation of human umbilical vein ECs and
TGF-β1 induces lncRNA NKILA expression to block NF-κB signaling, form a tube-like structure in vitro and in vivo in mice.424–426 This CSC
which inhibits metastasis of breast CSCs.412 TGF-β also directly plasticity has also been demonstrated in other tumors, including
regulates the expression of Wnt5a in breast CSCs to limit the stem neuroblastoma, renal, breast, and ovarian cancer.427–430
cell population.413 Furthermore, Notch, IKK/NF-κB, and other path- The vascular microenvironment maintains the initial undifferen-
ways together regulate the proliferation and metastasis of CD133+ tiated dormancy of stem cells, supports self-renewal, invasion and
cutaneous SCC stem cells.409 PI3K/mTOR signaling upregulates the metastasis of CSCs, and protects CSCs from any injury.431 The role of
expression of STAT3 to promote the survival and proliferation of the EC signaling system has been proven in maintaining the survival
breast CSCs.328 Inhibition of TORC1/2 increases FGF1 and Notch1 and self-renewal of head and neck SC stem cells.432 Pasquier and
expression. The PI3K/AKT/mTOR and Sonic Hh pathways cooperate colleagues433 showed that treatment with EC microparticles in breast
to inhibit the growth of pancreatic CSCs.414 Increasing evidence and ovarian cancer models increased the number of CSCs and
shows that crosstalk regulates the survival, self-renewal, and promoted sphere formation of CSCs. The interaction between CSCs
metastasis of CSCs. and blood vessels promotes the self-renewal of CSCs through the
VEGF-Nrp1 loop.418 CSCs promote cancer angiogenesis by inducing
The microenvironment of CSCs secretion of the cytokines VEGF and hepatocyte growth factor (HGF)
CSCs interact with the microenvironment through adhesion from ECs.434 VEGF receptor 2 plays a key role in vasculogenic mimicry
molecules and paracrine factors. The microenvironment provides formation, neovascularization, and tumor initiation of glioma stem-
a suitable space for the self-renewal and differentiation of CSCs, like cells.435 As a result, the secretion of VEGF in stem cell-like glioma
protects CSCs from genotoxicity, and increases their chemical and cells is higher than that in normal cancer cells424 and regulates the
radiological tolerance. The TME mainly consists of the tumor proliferation of glioma stem cells through the mTOR signaling
stroma, adjacent tissue cells, microvessels, immune cells, and pathway.436 Subsequent studies have further shown that multiple
immune molecules.415 CSCs not only adapt to changes in the TME signals, such as integrin, Notch, and growth factor receptors, are
but also affect the TME. Concurrently, the microenvironment also linked to each other on the cell surface to maintain the stemness of
promotes the self-renewal of CSCs, induces angiogenesis, recruits CSCs.437,438
immune and stromal cells, and promotes tumor invasion and
metastasis (Fig. 4). The hypoxia microenvironment and CSCs. Hypoxia is a key
component for CSC formation and maintenance.439 The hypoxic
Vascular niche microenvironments and CSCs. The normal vascu- microenvironment maintains the undifferentiated state of cancer
lature is composed of ECs, basement membranes, and parietal cells. cells, enhances their cloning rate, and induces the expression of
ECs are the basis for the formation of the inner surface of blood CD133 as a specific biomarker of CSCs.440 HIFs are important
vessels.416 Studies reported that glioblastoma stem cells are located transcription factors that regulate cellular hypoxia responsive-
around the blood vessels, and the concept of the cancer ness441 and inhibit cell apoptosis.442 As a heterodimer, HIF is
microvascular environment was first proposed. Calabrese et al.417 composed of HIFα and HIFβ.443 HIF-1α regulates the proliferation
demonstrated that direct contact between ECs and CSCs occurs in and fate of CSCs in medulloblastoma and glioblastoma multi-
brain tumors. CSCs are also found near ECs in other cancers, such as forme444 and activates the NF-κB pathway to promote CSC
papilloma and colorectal cancer.418,419 A study also showed that survival and tumorigenesis.445 HIF-2α maintains the survival and
CD133+/CD144− glioma stem cell-like cells differentiate into cancer phenotype of CSCs.446 HIFα also regulates the expression of the
cells and endothelial progenitor cells and finally into mature ECs.420 target genes GLUT1, GLUT3, LDHA, and PDK1. Thus, CSCs can
CSCs differentiate into cancer vascular stem cells/progenitor cells and adapt to a new method of cell energy metabolism and avoid
are directly involved in angiogenesis or form vasculogenic mimicry apoptosis caused by hypoxia.447

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
14
HIFs also regulate the stemness of CSCs. Previous studies have immune cells.476 It is not clear whether there are differences in the
shown that CSCs need to activate HIF-1α and HIF-2α to maintain functions of CAFs from different sources. CAFs affect cancer cell
their self-sustainability under hypoxic conditions448 and obtain growth through cell–cell interactions and the secretion of various
pluripotency by upregulating the Sox2 and Oct4 genes.440 More invasive molecules, such as cytokines, chemokines, and inflam-
importantly, activation of C-MYC by HIF-2α is necessary to ensure matory mediators.477–479
undifferentiated CSCs.449 The Wnt and Notch signaling pathways CAFs in the TME play an indispensable role in the generation
regulated by hypoxia and can induce the EMT, which promotes and maintenance of CSCs.480 CAFs transform cancer cells into
the stemness of CSCs and increases the invasiveness and CSCs.481 Studies have shown that CAFs promote the EMT and
resistance to radiotherapy and chemotherapy.450 HIF-1α binds enhance the expression of prostate CSC markers482 by secreting
the Notch ICD and enhances its transcriptional activity. In the IL-6 and IL-1β in breast cancer.483,484 CAFs also secrete TGF-β and
hypoxic microenvironment of glioma, both HIF-1α and HIF-2α activate related pathways to increase ZEB1 transcription, which
require the Notch signaling pathway to ensure the self-renewal stimulate lung cancer cells to undergo EMT and CSC transforma-
and undifferentiated status of CSCs.451 tion.485 CAFs secrete matrix metalloproteinases, which induce the
EMT and promote the growth of stem cell-specific components in
Tumor-associated macrophages and CSCs. Macrophages are an tumors.482 Paracrine interaction between CAFs and CSCs is critical
important component of the innate immune response and are a for maintaining the CSC niche of lung CSCs.486 Fibroblast-derived
group of cells with plasticity and heterogeneity.452 Infiltrating and CCL-2 regulates CSCs through gap activation, thus promoting the
inflammatory macrophages originate from the precursors of bone progression of tumors.487 CAFs and adipocytes also secrete leptin,
marrow mononuclear cells.453 These precursor cells infiltrate which increases the globulation rate of breast CSCs in vitro.488
various tissues from blood vessels and differentiate into different CAFs also regulate the proliferation of CSCs by other signaling
subtypes in different microenvironments. There are two subtypes pathways. For example, CAFs increase the secretion of CCL-2 to
of macrophages: the M1 and M2 phenotypes. The M1 phenotype activate the Notch1/STAT3 pathway, which increases the expres-
has anti-inflammatory and anti-tumor effects and secretes sion of stem cell markers and upregulates the globulation rate in
proinflammatory factors such as interleukin-1 (IL-1), IL-12, IL-23, breast cancer.489 CAFs regulate TIC plasticity in HCC through c-
TNF-α, chemokine (C-X-C motif) ligand 5 (CXCL5), CXCL9, and Met/FRA1/HEY1 signaling.490 CAFs secrete high levels of IL-6 to
CXCL10. M2 macrophages are generally considered to be the activate Notch signaling through STAT3 Tyr705 phosphorylation,
phenotype of tumor-associated macrophages (TAMs),454–456 have thus promoting the stem cell-like characteristics of HCC cells.491
immunosuppressive and angiogenesis-promoting effects, and are Similar studies have shown that CAF-derived exons enhance colon
considered to be a tumor-promoting cell type.456,457 M2 macro- stem cell resistance to 5-fluorouracil by activating the Wnt
phages secrete CCL17 (C-C chemokine ligand 17), CCL22, and signaling pathway.492
CCL24 and have low expression of IL-12 and high expression of IL-
10. Cytokines secreted by macrophages affect the proliferation, Cancer-associated MSCs and CSCs. MSCs have high self-renewal
tumorigenic transformation, or apoptosis of CSCs through various ability and multidirectional differentiation potential.493 MSCs also
signaling pathways.458 specifically migrate to the injured site and tumor tissue and are
TAMs are closely related to CSCs or stem cell transformation. easy to isolate and expand in vitro.494,495 MSCs are considered to
Renal epithelial cells cocultured with macrophages induce the be an ideal vector for gene therapy because of their characteristics
EMT to transform renal cancer cells into CSCs expressing CD117, of homing to and secreting cytokines in tumors.496 However,
Nanog, and CD133.459 Another study also showed that mucin-1 these tumorigenic characteristics of MSCs still need to be studied.
secreted by M2 macrophages induces the transdifferentiation of MSCs not only promote tumor development497,498 but also inhibit
non-small-cell lung cancer cells into CSCs that express CD133 and cancer cell growth.499 Bone marrow MSCs promote tumor growth
Sox2.460 Jinushi and colleagues461 also reported that TAMs secrete by promoting angiogenesis, metastasis, and the survival of
MFG-E8, which maintains the self-renewal ability of colon and CSCs.500 MSCs in the TME are conducive to the proliferation,
breast CSCs, and knockout of MFG-E8 significantly inhibits the carcinogenesis, and metastasis of breast CSCs through ionic
tumorigenic ability in SCID mice.461 TAMs are closely related to purinergic signal transduction.501 MSCs can differentiate into
glioma stem cell growth.462 TAMs are mainly distributed near CAFs, and CAFs further regulate CSCs and promote the occurrence
CD133+ glioma stem cells and accumulate in pericapillary and and metastasis of cancers.502 The possible mechanism is related to
hypoxic areas.463 Glioma stem cells recruit and maintain macro- the spontaneous fusion between cancer cells and MSCs.503 The
phages by secreting a potent chemokine membrane protein.464 fusion of MSCs with breast cancer, ovarian cancer, gastric cancer,
The ablation of TAMs inhibits the tumorigenesis of glioma stem and lung cancer cells in vitro and in vivo has been con-
cells.465 Recent studies have shown that the interaction between firmed.504,505 MSCs regulate the TME by secreting IL-6 to maintain
the TME and CSCs is regulated by a variety of signaling the undifferentiated state of osteosarcoma cells.506,507 IL-1
pathways.466 Macrophages enhance the invasion of glioma stimulates the secretion of PGE2 via autocrine signaling, which
stem-like cells through the TGF-β1 signaling pathway.467 TAMs ultimately activates β-catenin signaling in cancer cells in a
activate the STAT3/Sox2 signaling pathway in mouse breast CSCs paracrine manner and transforms cancer cells into CSCs.508 In
by secreting EGF, which promotes the self-renewal ability of the ECM, bone mesenchymal stem cells activate the NF-κB
CSCs.468 IL-8 secreted by TAMs also induces the EMT in pathway and induce a CSC phenotype by secreting a variety of
hepatocellular cancer cells by activating the JAK2/STAT3/Snail cytokines and chemokines, such as CXCL12, CXCL7, and IL-6/IL-
pathway.469 8.509 The interaction between MDSCs and CSCs via IL-6/STAT3 and
Notch signaling is critical to the progression of breast cancer.510
Cancer-associated fibroblasts and CSCs. CAFs are one of the most
important components of the TME and are critical in tumor Extracellular matrix and CSCs. The ECM is an insoluble structural
development and metastasis.470 The origin of these cells in the component of the matrix in mesenchymal and epithelial vessels.
stroma is not entirely clear. Current studies hypothesize that there The ECM includes collagen, elastin, aminoglycan, proteoglycan,
are five possible sources: (1) transference of fibroblasts in the host and noncollagen glycoprotein.511,512 At present, increasing
stroma;471 (2) EMT;472 (3) transdifferentiation of perivascular evidence shows that the ECM is an integral part of stem cell
cells;473 (4) EMT;474 and (5) differentiation of MSCs derived from niches that regulates the balance of stem cells in three different
bone marrow.475 In addition, CAFs are also derived from other cell biological states: static, self-renewal, and differentiation.513 Experi-
types, such as smooth muscle cells, pericytes, adipocytes, and ments in vitro and in vivo have shown that ECM receptors can be

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
15
used to aggregate CSCs514 and induce drug resistance.513,515 some serious adverse reactions still occur.549 Subsequently, to
Fibronectin, vimentin, collagen, and proteoglycan in the ECM bind improve the availability and affordability of radioimmunotherapy
to cytokines such as FGF, HGF, VGF, BMP, and TGF-β in the TME for refractory or recurrent non-Hodgkin’s lymphoma (NHL), a
and regulate their activities.516 In HCC, an increased matrix phase II clinical trial for a radioiodine replacement of rituximab
promotes cell proliferation and chemotherapeutic resistance and was carried out, which showed a response rate of 71% and a
increases the expression of CSC-related markers, including CD44, complete remission rate of 54% in 35 patients, with only two cases
CD133, c-kit, cxcr4, Oct4, and Nanog. Hyaluronic acid in the ECM is of grade IV hematologic toxicity observed.550 Encouragingly,
a ligand for the CD44 receptor and can regulate the acquisition alemtuzumab, a humanized CD52 antibody, has been approved
and maintenance of CSC stemness during mutual contact.517 The for the treatment of chronic lymphocytic leukemia (CLL) in
ECM also binds the Wnt ligand Wnt5b via molecular MMP3 and patients who failed to respond to alkylating agents and purine.
leads to the expansion and proliferation of mammary epithelial Furthermore, the combination of the CD20 and CD52 antibodies in
stem cells.518 In addition, tenascin C in the ECM maintains the the treatment of refractory CLL was safe, nontoxic, feasible, and
stability of breast CSCs by increasing the activity of the Wnt and positive.551 Another antibody drug, relabeled bivatuzumab, is an
Notch signaling pathways.519 anti-CD44v6 mAb,71 which was found to be safe when it was used
for the treatment of head and neck SCC.552 These results have
Exosomes in the TME and CSCs. Exosomes are nanovesicles been obtained in subsequent clinical research553 and safety/
secreted by various types of living cells (30–100 nm in efficacy studies.554 Unfortunately, in a stage I dose escalation
diameter)520 and are widely distributed in peripheral blood, saliva, study with the CD44v6 antibody, one patient with head and neck
urine, ascites, pleural effusion, breast milk, and other body SCC of the esophagus suffered deadly skin toxicity.555
fluids.521 Exosomes contain a large number of functional proteins, Several CD123 antibodies have been developed, XmAb14045
RNA, microRNAs, DNA fragments, and other bioactive sub- and MGD006, and were designed with biospecific effects
stances.522–525 These bioactive substances mediate material against CD3 and CD123. Talacotuzumab is also effective against
transport and information exchange between cells, thus affecting CD16 and CD123. CSL360, another CD123 antibody, was used to
the physiological function of cells.526,527 The exosomes secreted treat relapsed, refractory, or high-risk acute myeloid leukemia
by cancer cells promote angiogenesis,528 induce differentiation of (AML) and displayed no anti-leukemic activity in most cases.556
tumor-related fibroblasts,529 participate in immune regulation of SL-401, another CD123 antibody, was used to treat blastic
the TME,530 and regulate the microenvironment before metas- plasmacytoid dendritic cell neoplasm patients. The results
tasis.531 Clinical analysis has revealed that exosomes are released showed major positive responses in seven out of nine patients,
at higher levels in cancer cells.532 including five complete responses and two partial responses.557
Recent studies have shown that endocytosis of lipid rafts in MSCs An ongoing phase II study of SL-401 in 29 patients with blastic
is associated with increased secretion of exosomes.533 Exosome plasmacytoid dendritic cell neoplasms demonstrated robust
signaling mediates the interaction of CSCs and normal stem cells, single-agent activity with an 86% overall response rate.558 The
thereby regulating oncogenesis and tumor development.534 Exo- latest antibodies against CSC surface markers, such as
somes also regulate CSC growth by targeting specific signaling XmAb14045 (NCT02730312), flotetuzumab (NCT02152956),
pathways, such as Wnt, Notch, Hippo, Hh, and NF-κB.535–537 and talacotuzumab (NCT02472145), are also in clinical study.
Extracellular vesicles released by glioblastoma stem cells promote Furthermore, several other markers that can distinguish LSCs
neurosphere formation, endothelial tube formation, and the from other cells are under clinical development, such as IL-1
invasion of glioblastoma.538 CSCs promote cell proliferation and receptor accessory protein, CD27/70, CD33, CD38, CD138, CD93,
self-renewal through crosstalk between exosome signal transduction CD99, C-type lectin-like molecule-1, and T cell immunoglobulin
and the surrounding microenvironment.539 The exosomes released mucin-3.
from CSCs affect signal transduction in nearby breast cancer cells540 EpCAM, a common CSC biomarker, has also received attention
and increase the stemness of breast cancer cells.540 Similarly, in clinical trials.559 Adecatumumab, an EpCAM antibody, was used
fibroblast-derived exosomes contribute to chemoresistance by in patients with hormone-resistant prostate cancer, and the results
promoting colorectal CSC growth.491 Exosomes in the TME promote showed that the EpCAM-specific antibody has great clinical
the transformation of non-CSCs into CSCs. CAF-derived exosomes potential.560 Catumaxomab, a multifunctional mAb against
significantly increase the ability to form mammary globules and EpCAM, binds and recognizes EpCAM and the T cell antigen
promote the stemness of breast cancer cells.541 Similarly, CAF- CD3 (anti-EpCAM × anti-CD3).561 Intraperitoneal injection of catu-
derived exosomes also promote sphere formation of colorectal maxomab to treat EpCAM-positive ovarian cancer and malignant
cancer cells by activating Wnt signaling and ultimately increase the ascites has shown high efficacy in killing cancer cells and reducing
percentage of CSCs.491 Exosomes from glioma-associated MSCs the formation of ascites.562 Catumaxomab has been used in non-
increase the tumorigenicity of glioma stem-like cells by transferring small-cell lung cancer and also had a good survival rate.561
miR-1587.542 In addition, exosomes regenerate stem cell phenotypes However, other types of EpCAM antibodies, such as edrecolo-
by mediating the EMT or regulating stem cell-related signaling mab563 and adecatumumab,564 have minimal efficacy in colorectal
pathways, such as the Wnt pathway, Notch pathway, Hh pathway and breast cancers. Combining EpCAM antibodies with chimeric
and other pathways, which convert cells into CSCs.543 Exosomes antigen receptor T cell (CAR-T) technology has also been used in
have many advantages, such as low immunogenicity, biocompat- various types of cancers in phase I trials, such as NCT02915445,
ibility, easy production, cytotoxicity, easy storage, high drug loading NCT03563326, NCT02729493, and NCT02725125. With a deeper
capacity, and long life and have become ideal drug carriers for understanding of CSC surface biomarkers, there has been
cancer therapy.544–548 significant progress in developing antibodies targeting CSCs
(Table 2). However, CSC surface phenotypes can vary in different
patients or different cancers, and different CSC populations with
THERAPEUTIC TARGETING OF CSCS different phenotypes might coexist. CSCs also diverge or evolve
Agents targeting CSC-associated surface biomarkers in clinical into different cancer cells, acquiring distinct phenotypes upon
trials relapse. Therefore, the strategies used in clinical trials should be
Monoclonal antibodies (mAbs) that target CSC-specific surface determined according to the phenotypes of the different cancers.
biomarkers have become an emerging technology for cancer At the same time, combining different surface antibodies with
therapy. Rituximab, a CD20 mAb, is an active agent for the relevant chemotherapy drugs can achieve an ideal therapeutic
treatment of follicular lymphoma and mantle-cell lymphoma, but effect.

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
16
Table 2. Agents targeting CSC-associated surface markers in ongoing clinical trials

Drug name Antibody target Condition Sample size Highest status NCT number Current status

Surface antigens
Catumaxomabr (emovab) EpCAM/CD3 Ovarian cancer II 44 NCT00189345 Completed
Tagraxofusp CD123 Acute myeloid leukemia I 36 NCT03113643 Recruiting
SL-401
KHK2823 I 39 NCT02181699 Terminated
Talacotuzumab III 326 NCT02472145 Completed, has results
SGN-CD123A I 17 NCT02848248 Terminated
IMGN632 II 155 NCT03386513 Recruiting
XmAb14045 CD123/CD4 II 105 NCT02730312 Recruiting
MGD006 CD123/CD3 II 179 NCT02152956 Recruiting
JNJ-63709178 III 326 NCT02472145 Completed, has results
CSL362 CD124 I 30 NCT01632852 Completed
TTI-621 CD47 Solid tumor I 260 NCT02663518 Recruiting
Hu5F9-G4 Solid tumor I 88 NCT02216409 Completed
IBI188 Advanced malignancies I 42 NCT03763149 Recruiting
CC-90002 Hematologic neoplasms I 28 NCT02641002 Terminated
AO-176 Solid tumor I 90 NCT03834948 Recruiting
SRF231 Solid tumor I 148 NCT03512340 Recruiting
Bivatuzumab mertansine Metastatic breast cancer I 24 NCT02254005 Completed
Vadastuximab talirine (SGN-CD33A) CD33 Acute myelogenous leukemia I 195 NCT01902329 Completed
IMGN779 I 62 NCT02674763 Completed
Mylotarg (gemtuzumab ozogamicin) ECG IV 56 NCT03727750 Recruiting
RO5429083 CD44 Malignant solid tumors I 65 NCT01358903 Completed
SPL-108 Ovarian cancer I 18 NCT03078400 Recruiting
Salazosulfapyridine CD44V4 Non-small-cell lung cancer I UMIN000017854
AMC303 CD44V6 Solid tumor I 55 NCT03009214 Recruiting
Immune checkpoints
Ipilimumab CTLA-4 Non-small-cell lung cancer II 24 NCT01820754 Completed, has results
Nivolumab PD-1 Glioblastoma multiforme II 29 NCT02550249 Completed
Pembrolizumab II 80 NCT02337491 Completed, has results
Cemiplimab II 30 NCT04006119 Recruiting
Idarubicin Acute myeloid leukemia II 51 NCT01035502 Completed
Sym021 Solid tumor lymphomas I 102 NCT03311412 Recruiting
Durvalumab Solid tumors II 124 NCT02403271 Completed, has results
Atezolizumab PD-L1 Non-small-cell lung cancer III 1225 NCT02008227 Completed, has results
Avelumab Recurrent glioblastoma II 52 NCT03291314 Completed
Sym023 Tim3 Solid tumor I 48 NCT03489343 Recruiting
ARGX-110 CD70 Acute myeloid leukemia II 36 NCT03030612 Active, not recruiting
Varlilumab (CDX-1127) Solid tumors II 175 NCT02335918 Completed
Sym022 LAG3 Solid tumor I 30 NCT03489369 Recruiting
MGD013 CD70/LAG3 Solid tumors I 255 NCT03219268 Recruiting

Agents targeting CSC-associated signaling pathways in clinical pathways. For example, GSIs have been tested in clinical trials.
trials Among them, MK-0752 (NCT00100152) was the first GSI used to
The signaling pathways that regulate the maintenance and treat T cell acute lymphoblastic leukemia in children in a phase I
survival of CSCs have become targets for cancer treatment. At trial. However, the study was terminated because of poor
present, the main signaling pathways are the Wnt, Notch, and Hh results.575 MK-0752 also had no clinical activity in extracranial
signaling pathways, as well as the TGF-β, JAK-STAT, PI3K, and NF- solid tumors in subsequent phase II trials. Only one complete
κB signaling pathways. These pathways often interact with each response with interdegenerative astrocytoma and SD extension
other during tumor development and in CSCs. Considerable out of 10 patients with different types of glioma was
progress has been made in early clinical trials for Notch and Hh observed.576 MK-0752 is well tolerated and shows targeted
pathway inhibitors, while targeting the Wnt pathway has proven inhibition in recurrent pediatric central nervous system
to be difficult.10 tumors.577 In addition, combining MK-0752 with cisplatin
The Notch signaling pathway plays an important role in the treatment for ovarian cancer,578,579 docetaxel treatment for
maintenance of CSCs565,566 and can induce CSC differentiation. locally advanced or metastatic breast cancer,569 and gemcita-
Abnormal activity of the Notch signaling pathway has been bine treatment for ductal adenocarcinoma of the pancreas580
observed in many cancers, such as leukemia,567 glioblas- has shown good efficacy. However, the clinical effect was
toma,568,569 breast cancer,570 lung cancer,571 ovarian cancer,572 minimal in patients with advanced solid tumors,576,581 including
pancreatic cancer,573 and colon cancer.574 At present, there are metastatic pancreatic cancer.582
three major clinical methods used to inhibit Notch signaling, In addition, RO4929097, another selective GSI, showed good
secretase inhibition (γ-secretase inhibitor (GSI)), Notch receptor anti-tumor activity in preclinical and early trials,583,584 but was not
or ligand antibodies, and combination therapy with other good for metastatic colorectal cancer,585 metastatic pancreatic

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
17
adenocarcinoma,586 or recurrent platinum-resistant ovarian can- treatments in phase I/II clinical trials, such as NCT02111187 for
cer.587 Combinations of RO4929097 with gemcitabine,588 temsir- prostate cancer, NCT02027376 for breast cancer, and
olimus,587 cediranib,589 or capecitabine590 in advanced solid NCT02195973 for recurrent ovarian cancer.
tumors, as well as with bevacizumab in recurrent high-grade Glasdegib was the first Hh pathway inhibitor approved for the
glioma, are well tolerated and have modest clinical benefits. treatment of AML in patients older than 75 years or those unable
However, NCT01154452, the combination of RO4929097 with to use intensive induction chemotherapy601 and showed good
vismodegib and vismodegib alone for patients with advanced safety and tolerability in a phase I trial for patients with partial
osteosarcoma, showed no significant difference in a phase Ib trial. hematologic malignancies in Japan.613 In a phase II trial, glasdegib
The third oral GSI, PF-03084014, had good efficacy in desmoid combined with cytarabine/daunorubicin had a significant efficacy
tumors either in phase I or subsequent phase II studies.591 in patients with AML, chronic myeloid leukemia (CML) or high-risk
Preliminary evidence of its clinical efficacy was demonstrated in myelodysplastic syndromes.614 Glasdegib combined with low-
patients with solid tumors,592 as well as in patients with recurrent dose cytarabine (LDAC) is a potential option for AML patients who
acute T cell lymphoblastic leukemia.593 Other selective GSIs, such are not suitable for intensive chemotherapy.615 Other selective
as BMS-906024 (NCT01292655), BMS-986115 (NCT01986218), CB- SMO inhibitors, including taladegib (LY2940680) and saridegib
103 (NCT03422679), LY3039478 (NCT02836600), and LY900009 (IPI-926), have also entered clinical trials for other cancers. As
(NCT01158404), have also entered the clinical trial stage, and the single-target agents, these SMO inhibitors have drug resistance
results still need to be verified. problems. To reduce this problem, some novel inhibitors of
DLL4 plays a vital role in regulating tumor angiogenesis.594 terminal components of Hh signaling pathway are being
Therefore, targeting DLL4 is another strategy to block Notch developed, such as arsenic trioxide (ATO)616 and GANT-61.617
signaling, and this is being tested in the clinic. Demcizumab (OMP- The Wnt signaling pathway is associated with tumor develop-
21M18), a humanized IgG2 mAb that targets DLL4 and blocks its ment in breast cancer,618 ovarian cancer,619 esophageal squamous
interactions with Notch receptors, was tested in a phase I dose cell cancer,620 colon cancer,621 prostate cancer,622 and lung
escalation study with 55 patients with previously treated solid cancer.623 Until now, several drugs aimed at the Wnt signaling
tumors.595 The results have shown that demcizumab had good pathway have been in clinical trials, while the majority of Wnt
efficacy against solid tumors, but was not good for metastatic inhibitors are in preclinical testing. Clinical data from initial trials
pancreatic cancer treatment when combined with gemcitabine have shown that ipafricept (OMP-54f28/FZD8-Fc) is a first-in-class
and Abraxane (NCT02289898). NCT02259582, a combination of recombinant fusion protein that antagonizes Wnt signaling.624
demcizumab with carboplatin and pemetrexed to treat lung However, its role in patients with desmoid cancers and germ cell
cancers (DENALI study), is ongoing in another phase II study.595 cancers is negligible.625 NCT02050178, ipafricept combined with
Enoticumab, another fully human IgG1 antibody against DLL4, has ab-paclitaxel and gemcitabine in patients with untreated stage IV
promising activity in phase I clinical trials for advanced solid pancreatic cancer, NCT02092363, ipafricept combined with
malignancies. paclitaxel and carboplatin in patients with recurrent platinum-
Activation of Hh signaling has been implicated in a variety of sensitive ovarian cancer, and NCT02069145, ipafricept combined
cancers.596–598 Activation of Hh signaling in CSCs contributes to with sorafenib in patients with HCC, are currently being
CSC stemness, chemoresistance, and metastatic dissemination. investigated. PRI-724, a β-catenin inhibitor, inhibits the interaction
The Hh signaling pathway mainly regulates target gene expression between β-catenin and its transcriptional coactivators. Safety and
via smoothened (SMO)-mediated nuclear transfer of transcription efficacy testing of PRI-724 for patients with advanced myeloid
factors. Three oral SMO antagonists, vismodegib (GDC-0449), malignancies (NCT01606579) and advanced or metastatic pan-
sonidegib (LDE225), and glasdegib (PF-04449913), have been creatic cancer (NCT01764477) have been completed in phase I
approved by the Food and Drug Administration (FDA) and show studies. CWP232291, another inhibitor of β-catenin activity, has
significant activity in locally advanced and metastatic basal cell also been shown to be effective for AML (NCT03055286) in a
carcinoma, as well as in AML.599–601 Vismodegib was the first phase I clinical study and for recurrent or refractory myeloma
proposed Hh pathway inhibitor in cancer research602 and is (NCT02426723) in a phase I/II clinical study.626 Other Wnt signaling
approved by the FDA603 for local or advanced metastatic basal cell inhibitors have also been under clinical trial, including LGK974
carcinoma treatment.599 Subsequently, phase I and phase II trials (NCT02278133), ETC-159 (NCT02521844), and OMP-18R5
targeting recurrent medulloblastoma have shown that the (NCT01973309, NCT01957007, and NCT02005315).
progression-free survival (PFS) of Shh-mb patients treated with In addition, the mitochondrial glycolysis pathway also plays a
vismodegib is longer and more effective than that of non-Shh-mb key role in regulating the proliferation and apoptosis of CSCs.
patients. Vismodegib even has better activity in patients with Venetoclax, a BCL-2 inhibitor, was initially approved by the FDA
recurrent Shh-mb but not in patients with recurrent non-Shh- recently and shows good tolerance and activity for AML patients
mb.604,605 Vismodegib has also been tested in metastatic color- with adverse reactions.627 Two arachidonate 5-lipoxygenase
ectal cancer,606 pancreatic cancer,607 chondrosarcoma,608 inhibitors, VIA-2291 and GSK2190915, might be potent agents
relapsed/refractory NHL, CLL,609 and ovarian cancer.610 Disap- for targeting LSCs in CML,628 as shown in Table 3.
pointingly, these treatments with vismodegib have not resulted in Other abnormal signaling pathways have also been found
better survival. in CSCs, such as TGF-β, JAK-STAT, PI3K, and NF-κB. These
Sonidegib was the second SMO antagonist approved for the signaling pathways are not independent of each other but
treatment of locally advanced basal cell carcinoma that recurred rather form a complex signaling network. Agents targeting CSC-
after surgery or radiotherapy and is not suitable for surgery or associated signaling pathways in ongoing clinical trials are
radiation therapy.611 In addition, the results of a multicenter, listed in Table 3.
randomized, double-blind phase II trial have shown that 200 mg
sonidegib for patients with advanced basal cell carcinoma is the Targeting the CSC microenvironment
most clinically appropriate dose.600 The CSC microenvironment contributes to the self-renewal and
In a phase I study of a 3 + 3 dose escalation to treat small-cell differentiation of CSCs and regulates CSC fate by providing cues in
lung cancer patients, sonidegib combined with cisplatin and the form of secreted factors and cell contact. CXCR4 has been
etoposide sustained PFS in patients with Sox2 amplification.224 found in most cancers, especially in CSCs. The most well-
These combinations in a phase II trial for patients with recurrent characterized drug-targeting CXCR4 is plerixafor (AMD3100), and
medulloblastoma resulted in a complete or partial response in this drug is an effective hematopoietic stem cell mobilizer for
50% of patients612 and have been used for other cancer patients with multiple myeloma and NHL.629 AMD3100 treatment

Signal Transduction and Targeted Therapy (2020)5:8


18
Table 3. Agents targeting CSC-associated signaling pathways and microenvironment in ongoing clinical trials

Drug name Target Condition Phase Sample size NCT number Current status

Hedgehog inhibitors
Vismodegib (GDC-0449) Smoothened Recurrent or refractory medulloblastoma II 31 NCT00939484 Completed, has results
Basal cell carcinoma 28 NCT01700049 Completed, has results
Sarcoma 78 NCT01700049 Completed, has results
Recurrent small-cell lung carcinoma 168 NCT01700049 Completed, has results
Metastatic pancreatic cancer 98 NCT01088815 Completed, has results
Ovarian cancer 104 NCT00739661 Completed, has results
Metastatic colorectal cancer 199 NCT00636610 Completed, has results
Sonidegib (LDE225) Basal cell carcinoma II 10 NCT01350115 Completed, has results
Relapsed medulloblastoma 20 NCT01708174 Completed, has results
Acute myeloid leukemia 70 NCT01826214 Completed, has results
Pancreatic adenocarcinoma 20 NCT01431794 Completed, has results
Advanced or metastatic hepatocellular carcinoma 9 NCT02151864 Completed
Recurrent plasma cell myeloma 28 NCT02086552 Active, not recruiting,
has results
Advanced pancreatic cancer 39 NCT01485744 Active, not recruiting
Advanced breast cancer I 12 NCT02027376 Completed, has results
Glasdegib Acute myeloid leukemia II 255 NCT01546038 Completed, has results
BMS-833923 (XL139) Solid tumors II 12 NCT01413906 Completed
Small-cell lung carcinoma 5 NCT00927875 Completed
Yang et al.
Targeting cancer stem cell pathways for cancer therapy

Metastatic gastric, gastroesophageal, esophageal 39 NCT00909402 Completed


adenocarcinomas
Advanced or metastatic basal cell carcinoma 53 NCT00670189 Completed
Leukemia 70 NCT01357655 Terminated, has results
Taladegib (LY2940680) Localized esophageal or gastroesophageal II 9 NCT02530437 Active, not recruiting
junction cancer
Small-cell lung carcinoma 26 NCT01722292 Terminated, has results
LEQ-506 Solid tumors I 57 NCT01106508 Completed
G-024856 BCC I
Patidegib (IPI-926) Basal cell carcinomas II 36 NCT02828111 Completed, has results
Metastatic or locally advanced chondrosarcoma 105 NCT01310816 Completed
Metastatic pancreatic cancer 122 NCT01130142 Completed
Recurrent head and neck cancer I 9 NCT01255800 Completed
Notch inhibitors
MK-0752 γ-Secretase Advanced breast cancer I 103 NCT00106145 Completed
Pancreatic cancer I 44 NCT01098344 Completed
Metastatic breast cancer I/II 30 NCT00645333 Completed, has results
RO4929097 Recurrent melanoma II 14 NCT01196416 Completed, has results
Advanced or metastatic sarcoma 78 NCT01154452 Completed, has results
Recurrent renal cell carcinoma 12 NCT01141569 Completed, has results

Signal Transduction and Targeted Therapy (2020)5:8


Advanced solid tumors 20 NCT01131234 Completed
Recurrent and/or metastatic epithelial ovarian cancer, 45 NCT01175343 Completed, has results
fallopian tube cancer, or primary peritoneal cancer
Metastatic pancreas cancer 18 NCT01232829 Completed, has results
Recurrent colon cancer 37 NCT01116687 Completed, has results
Table 3 continued
Drug name Target Condition Phase Sample size NCT number Current status

Recurrent or refractory non-small-cell lung cancer 7 NCT01070927 Completed


Nirogacestat (PF-03084014) Metastatic cancer pancreas II 3 NCT02109445 Terminated, has results
Fibromatosis II 17 NCT01981551 Active, not recruiting
Triple-negative breast neoplasms II 19 NCT02299635 Terminated, has results
LY900009 Advanced cancer I 35 NCT01158404 Completed, has results
Crenigacestat (LY3039478) Pan-Notch Advanced solid tumor I 12 NCT02836600 Active, not recruiting
T cell acute lymphoblastic leukemia, T cell I/II 36 NCT02518113 Completed, has results
lymphoblastic lymphoma
AL101 Adenoid cystic carcinoma II 36 NCT03691207 Recruiting
CB-103 Advanced or metastatic solid tumors and I/II 165 NCT03422679 Recruiting
hematological malignancies
BMS-906024 Advanced or metastatic solid tumors I 94 NCT01292655 Completed
Lymphoblastic leukemia, acute T cell I 31 NCT01363817 Completed
Demcizumab (OMP-21M18) DLL4 Pancreatic cancer II 207 NCT02289898 Completed, has results

Signal Transduction and Targeted Therapy (2020)5:8


Non-squamous, non-small-cell neoplasm of lung II 82 NCT02259582 Completed, has results
Brontictuzumab (OMP-52M51) Adenoid cystic carcinoma Not applicable 1 NCT02662608 Completed, has results
Enoticumab (MEDI528) Advanced solid malignancies I 83 NCT00871559 Completed
MEDI0639 Solid tumors I 58 NCT01577745 Completed, has results
Wnt inhibitors
Ipafricept (OMP-54F28) Wnt receptor Solid tumors I 26 NCT01608867 Completed
Pancreatic cancer I 26 NCT02050178 Completed
Ovarian cancer I 37 NCT02092363 Completed
Hepatocellular cancer I 10 NCT02069145 Completed
Vantictumab (OMP-18R5) Metastatic breast cancer I 37 NCT01973309 Completed
Solid tumors I 35 NCT01345201 Completed
Pancreatic cancer I 30 NCT02005315 Completed
Yang et al.

PRI-724 β-Catenin/CBP Colorectal adenocarcinoma II 0 NCT02413853 Withdrawn


Acute myeloid leukemia 49 NCT01606579 Completed
Solid tumors 23 NCT01302405 Terminated
Advanced pancreatic cancer 20 NCT01764477 Completed
CWP232291 Acute myeloid leukemia I 69 NCT01398462 Completed
Multiple myeloma I 25 NCT02426723 Completed
LGK974 Porcupine Metastatic colorectal cancer I 20 NCT02278133 Completed
Pancreatic cancer I 170 NCT01351103 Recruiting
ETC-1922159 Solid tumors I 65 NCT02521844 Active, not recruiting
Other signaling pathways inhibitors
Galunisertib (LY2157299) TGF-β Prostate cancer II 60 NCT02452008 Recruiting
LY3200882 Colorectal cancer II 31 NCT04031872 Not yet recruiting
AVID200 Malignant solid tumor I 36 NCT03834662 Recruiting
Trabedersen (AP 12009) Pancreatic neoplasms II 62 NCT00844064 Completed
Breast cancer 16 NCT01959490 Completed, has results
Targeting cancer stem cell pathways for cancer therapy

Glioblastoma 141 NCT00431561 Completed


Fresolimumab (GC1008) Non-small-cell lung carcinoma II 60 NCT02581787 Recruiting
Metastatic Breast Cancer 23 NCT01401062 Completed, has results
Carcinoma 29 NCT00356460 Completed
Renal cell
Melanoma
19
Table 3 continued 20

Drug name Target Condition Phase Sample size NCT number Current status

Vactosertib (TEW-7197) Advanced-stage solid tumors I 35 NCT02160106 Completed


NIS793 Breast cancer I 220 NCT02947165 Recruiting
Lung cancer
Hepatocellular cancer
Ruxolitinib JAK Metastatic breast cancer III 29 NCT01594216 Completed
Myeloproliferative neoplasms 309 NCT00952289 Completed, has results
AZD4205 Advanced non-small-cell lung cancer II 120 NCT03450330 Recruiting
SAR302503 Hematopoietic neoplasm II 97 NCT01523171 Completed
SB1518 JAK/FLT3 Acute myelogenous leukemia II 76 NCT00719836 Completed
PI3K inhibitors
Alpelisib PI3K Advanced breast cancer II 90 NCT03386162 Recruiting
Buparlisib (BKM120) Triple-negative metastatic breast cancer II 50 NCT01629615 Completed
BYL719 Advanced or metastatic gastric cancer I 18 NCT01613950 Completed
SF1126 Advanced or metastatic solid tumors I 44 NCT00907205 Completed
SAR245409 PI3K and mTOR Advanced or metastatic solid tumors I 146 NCT01390818 Completed, has results
EGFR inhibitors
Bevacizumab EGFR Breast cancer I 75 NCT01190345 Completed
Matuzumab (EMD 72000) Esophageal cancer II 72 NCT00215644 Completed, has results
Non-small-cell lung carcinoma 150 NCT00111839 Completed, has results
Metabolism inhibitors
Venetoclax (ABT-199) BCL-2 Acute myelogenous leukemia II 32 NCT01994837 Completed, has results
Yang et al.
Targeting cancer stem cell pathways for cancer therapy

Pegzilarginase Recombinant pegylated Small-cell lung cancer II 84 NCT03371979 Active, not recruiting
arginase
131I-TLX-101 LAT1 Glioblastoma multiforme II 44 NCT03849105 Recruiting
Rifampicin FAS Advanced solid tumors I 36 NCT03077607 Completed, has results
TVB-2640 Advanced breast cancer II 80 NCT03179904 Recruiting
IM156 AMPK Advanced solid tumor I 36 NCT03272256 Recruiting
Telaglenastat Glutaminase Solid tumors II 85 NCT03965845 Recruiting
CB-1158 Arginase Advanced solid tumors II 5 NCT03361228 Completed
Niche inhibitors
Plerixafor (Mozobil) CXCR4 Advanced pancreatic, ovarian, and colorectal cancers I 26 NCT02179970 Completed
BL-8040 Metastatic pancreatic adenocarcinoma II 23 NCT02907099 Active, not recruiting
BKT140 Multiple myeloma II 16 NCT01010880 Completed
BMS-936564 Relapsed/refractory multiple myeloma I 46 NCT01359657 Completed
BMS-936564 Acute myelogenous leukemia I 98 NCT01120457 Completed
LY2510924 Solid tumor I 9 NCT02737072 Terminated, has results
MSX-122 Refractory metastatic or locally advanced solid tumors I 27 NCT00591682 Suspended
USL311 Advanced solid tumors and relapsed/recurrent II 120 NCT02765165 Recruiting
Glioblastoma multiforme
AMD3100 Acute myeloid leukemia II 52 NCT00512252 Completed, has results
Reparixin CXCR1/2 Breast cancer II 20 NCT01861054 Terminated

Signal Transduction and Targeted Therapy (2020)5:8


Defactinib (VS-6063) FAK Non-small-cell lung cancer II 55 NCT01951690 Completed
Targeting cancer stem cell pathways for cancer therapy
Yang et al.
21
for relapsed or refractory AML resulted in 46% of patients with used for CSC-directed immunotherapy, and most of them are
complete remission with or without white count recovery in a recruited.
phase I/II study.630 In addition, plerixafor with high-dose
cytarabine and etoposide treatment for children with relapsed
or refractory acute leukemia or myelodysplasia syndrome resulted CONCLUSIONS AND PERSPECTIVES
in two patients with complete remission and one patient with We can conclude that CSCs are a small population of cancer cells
incomplete hematologic recovery out of 18 patients in a phase I that have self-renewal capacity and differentiation potential,
study.631 LY2510924, a small cyclic peptide, is a potent and thereby conferring tumor relapse, metastasis,643 heterogeneity,644
selective antagonist of CXCR4 and is well tolerated with no serious multidrug resistance,645,646 and radiation resistance.647 Several
adverse events in a phase I trial.632 However, the combination of pluripotent transcription factors, including Oct4, Sox2, Nanog,
LY2510924 with sunitinib for patients with metastatic renal cell KLF4, and MYC and some intracellular signaling pathways,
carcinoma has no better effect than sunitinib alone in a phase II including Wnt, NF-κB, Notch, Hh, JAK-STAT, PI3K/AKT/mTOR,
trial.633 The combination of LY2510924 with carboplatin/etoposide TGF/Smad, and PPAR, as well as extracellular factors, including
for patients with extensive small-cell lung cancer also had no vascular niches, hypoxia, TAM, CAF, cancer-associated MSCs, the
significant effect compared with that of carboplatin/etoposide ECM, and exosomes, are essential regulators of CSCs. Drugs,
alone in a phase II study.634 The combination of LY2510924 with vaccines, antibodies, and CAR-T cells targeting these pathways
other drugs for gliomas (NCT03746080, NCT01977677, and have also been developed to target CSCs.648 Importantly, many
NCT01288573) and multiple myeloma (NCT00103662, NCT0122 clinical trials on CSCs have also been performed and show a
0375, and NCT00903968) is also under clinical trial. promising future for cancer therapy.
The microenvironment plays an important role in CSC growth, However, there are also multiple hurdles that need to be solved
and it is also a promising target for treatment. Agents targeting to effectively eliminate CSCs. First, the characteristics of many
the microenvironment in ongoing clinical trials are listed in CSCs in specific types of tumors are not well identified.649 Second,
Table 3. since most studies on CSCs are performed in immune-deficient
mice in the absence of an adaptive immune system, these models
CSC-directed immunotherapy do not recapitulate the biological complexity of tumors in the
In the early twentieth century, Paul Ehrlich first proposed the idea clinic.650 Third, CSCs exist in a specific niche that sustains their
that an intact immune system suppresses tumor development survival. However, isolated CSCs are used in most current studies
(advancing cancer therapy with present and Emerging Immuno- that lacks a microenvironment.651 Fourth, the environmental
Oncology Approaches). Based on the understanding of cellular factors in CSC niches are not well understood, and the relationship
immune regulation, new methods for cancer treatment have between TAMs/CAFs and CSCs has not been well studied.645 Fifth,
emerged. In addition to the antibodies against the CSC molecules since CSCs also share some signaling pathways with normal stem
mentioned above, some novel anti-CSC immunotherapeutic cells, not all the regulatory factors that contribute to CSCs are
approaches, such as immunologic checkpoint blocking or CAR-T appropriate for use as therapeutic targets in cancer treatment.
cell therapies, have been developed. Some drugs that target the Sixth, whether CSCs should be activated or arrested is an open
immune checkpoint receptors CTLA-4,635 PD-1 (nivolumab,636 question in cancer therapy.652 Seventh, novel signaling and more
pembrolizumab,637 and cemiplimab,638) and PD-L1 (avelumab,639 regulatory levels, such as RNA editing,653 epigenetics,654 and
durvalumab,640 and atezolizumab641) have also been in clinical cellular metabolism,655 should be considered in cancer therapy
trials. I ipilimumab, a CTLA-4 antibody, is approved by the FDA, because they also contribute to the stemness of CSCs. Eighth,
and initial clinical results showed good effectiveness in patients some inhibitors that target CSC signaling are not very specific, and
with metastatic melanoma.642 For CAR-T cell therapy, as shown in so new inhibitors need to be designed.656 Ninth, natural products
Table 4, CD19, CD20, CD22, CD123, EpCAM, and ALDH have been that target CSCs should also be studied in the future.657 Finally,

Table 4. CSC-directed immunotherapy in ongoing clinical trials

Trial description Condition Sample size Phase NCT Number Current status

CD19 CAR-T B cell leukemia and lymphoma II 80 NCT03398967 Recruiting


CD123 CAR-T CD122+ myeloid malignancies II 45 NCT02937103 Recruiting
CD22 CAR-T Recurrent or refractory B cell malignancy I/II 45 NCT02794961 Unknown
CD22 CAR-T B-ALL I 15 NCT02650414 Recruiting
CD33 CAR-T Myeloid malignancies I/II 45 NCT02958397 Recruiting
CD33 CAR-T CD32+ acute myeloid leukemia I 11 NCT03126864 Active, not recruiting
CD38 CAR-T B-ALL II 80 NCT03754764 Recruiting
CD138 CAR-T Multiple myeloma II 10 NCT03196414 Recruiting
MUC1 CAR-T/PD-1 KO Advanced esophageal cancer I/II 20 NCT03706326 Recruiting
EGFR IL-12 CAR-T Metastatic colorectal cancer I 20 NCT03542799 Not yet recruiting
MESO CAR-T Refractory–relapsed ovarian cancer I/II 20 NCT03916679 Recruiting
MESO-19 CAR-T Metastatic pancreatic cancer I 4 NCT02465983 Completed
LeY CAR-T Myeloid malignancies I/II 445 NCT02958384 Recruiting
MOv19-BBz CAR -T Recurrent high-grade serous ovarian cancer I 18 NCT03585764 Recruiting
LeY CAR-T Advanced cancer I 30 NCT03851146 Recruiting
EpCAM CAR-T Recurrent breast cancer I 30 NCT02915445 Recruiting
BCMA CAR-T Multiple myeloma II 80 NCT03767751 Recruiting

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
22
novel ways of targeting the microenvironment of CSCs are also 27. Le, N. H., Franken, P. & Fodde, R. Tumour–stroma interactions in colorectal
promising and need to be explored. cancer: converging on beta-catenin activation and cancer stemness. Br. J. Cancer
98, 1886–1893 (2008).
28. Mani, S. A. et al. The epithelial–mesenchymal transition generates cells with
properties of stem cells. Cell 133, 704–715 (2008).
ACKNOWLEDGEMENTS
29. Al-Hajj, M., Wicha, M. S., Benito-Hernandez, A., Morrison, S. J. & Clarke, M. F.
This work was supported by the National Key Research and Development Program of
Prospective identification of tumorigenic breast cancer cells. Proc. Natl Acad. Sci.
China (nos. 2016YFC1302204 and 2017YFC1308600), the National Science Founda-
USA 100, 3983–3988 (2003).
tion of China (nos. 81672502, 81872071, and 81902664) and the Natural Science
30. Hermann, P. C. et al. Distinct populations of cancer stem cells determine tumor
Foundation of Chongqing (no. cstc2019jcyj-zdxmX0033).
growth and metastatic activity in human pancreatic cancer. Cell Stem Cell 1,
313–323 (2007).
31. Collins, A. T., Berry, P. A., Hyde, C., Stower, M. J. & Maitland, N. J. Prospective
ADDITIONAL INFORMATION identification of tumorigenic prostate cancer stem cells. Cancer Res. 65,
Competing interests: The authors declare no competing interests. 10946–10951 (2005).
32. Garcia-Mayea, Y., Mir, C., Masson, F., Paciucci, R. & ME, L. L. Insights into new
mechanisms and models of cancer stem cell multidrug resistance. Semin. Cancer
REFERENCES Biol. S1044-579X (2019).
1. Siegel, R. L., Miller, K. D. & Jemal, A. Cancer statistics. CA Cancer J. Clin. 69, 7–34 33. Gottesman, M. M., Fojo, T. & Bates, S. E. Multidrug resistance in cancer: role of
(2019). ATP-dependent transporters. Nat. Rev. Cancer 2, 48–58 (2002).
2. Sun, Y. Translational horizons in the tumor microenvironment: harnessing 34. Ginestier, C. et al. ALDH1 is a marker of normal and malignant human mammary
breakthroughs and targeting cures. Med. Res. Rev. 35, 408–436 (2015). stem cells and a predictor of poor clinical outcome. Cell Stem Cell 1, 555–567
3. Batlle, E. & Clevers, H. Cancer stem cells revisited. Nat. Med. 23, 1124–1134 (2007).
(2017). 35. Singh, S. et al. Aldehyde dehydrogenases in cellular responses to oxidative/
4. Reya, T., Morrison, S. J., Clarke, M. F. & Weissman, I. L. Jn Stem cells, cancer, and electrophilic stress. Free Radic. Biol. Med. 56, 89–101 (2013).
cancer stem cells. Nature 414, 105 (2001). 36. Diehn, M. et al. Association of reactive oxygen species levels and radioresistance
5. Chen, W., Dong, J., Haiech, J., Kilhoffer, M. C. & Zeniou, M. Cancer stem cell in cancer stem cells. Nature 458, 780–783 (2009).
quiescence and plasticity as major challenges in cancer therapy. Stem Cells Int. 37. Cojoc, M., Mabert, K., Muders, M. H. & Dubrovska, A. A role for cancer stem cells
2016, 1740936 (2016). in therapy resistance: cellular and molecular mechanisms. Semin. Cancer Biol. 31,
6. Lapidot, T. et al. A cell initiating human acute myeloid leukaemia after trans- 16–27 (2015).
plantation into SCID mice. Nature 367, 645–648 (1994). 38. Tang, D. G. Understanding cancer stem cell heterogeneity and plasticity. Cell Res.
7. Bonnet, D. & Dick, J. E. Human acute myeloid leukemia is organized as a hier- 22, 457–472 (2012).
archy that originates from a primitive hematopoietic cell. Nat. Med. 3, 730–737 39. Hirschmann-Jax, C. et al. A distinct “side population” of cells with high drug
(1997). efflux capacity in human tumor cells. Proc. Natl Acad. Sci. USA 101, 14228–14233
8. Shimokawa, M. et al. Visualization and targeting of LGR5(+) human colon cancer (2004).
stem cells. Nature 545, 187–192 (2017). 40. Quintana, E. et al. Phenotypic heterogeneity among tumorigenic melanoma
9. Shibata, M. & Hoque, M. O. Targeting cancer stem cells: a strategy for effective cells from patients that is reversible and not hierarchically organized. Cancer Cell
eradication of cancer. Cancers 11, 732 (2019). 18, 510–523 (2010).
10. Visvader, J. E. & Lindeman, G. J. Cancer stem cells: current status and evolving 41. Singh, S. K. et al. Identification of human brain tumour initiating cells. Nature
complexities. Cell Stem Cell 10, 717–728 (2012). 432, 396–401 (2004).
11. Ajani, J. A., Song, S., Hochster, H. S. & Steinberg, I. B. Cancer stem cells: the 42. van den Hoogen, C. et al. High aldehyde dehydrogenase activity identifies
promise and the potential. Semin. Oncol. 42(Suppl. 1), S3–S17 (2015). tumor-initiating and metastasis-initiating cells in human prostate cancer. Cancer
12. Bjerkvig, R., Tysnes, B. B., Aboody, K. S., Najbauer, J. & Terzis, A. J. Opinion: the Res. 70, 5163–5173 (2010).
origin of the cancer stem cell: current controversies and new insights. Nat. Rev. 43. Zhang, W. C. et al. Glycine decarboxylase activity drives non-small cell lung
Cancer 5, 899–904 (2005). cancer tumor-initiating cells and tumorigenesis. Cell 148, 259–272 (2012).
13. Matsui, W. et al. Characterization of clonogenic multiple myeloma cells. Blood 44. Miltenyi, S., Muller, W., Weichel, W. & Radbruch, A. High gradient magnetic cell
103, 2332–2336 (2004). separation with MACS. Cytometry 11, 231–238 (1990).
14. Hill, R. P. Identifying cancer stem cells in solid tumors: case not proven. Cancer 45. de Wynter, E. A. et al. Comparison of purity and enrichment of CD34+ cells from
Res. 66, 1891–1895 (2006). discussion 1890. bone marrow, umbilical cord and peripheral blood (primed for apheresis) using
15. Huntly, B. J. & Gilliland, D. G. Leukaemia stem cells and the evolution of cancer- five separation systems. Stem Cells 13, 524–532 (1995).
stem-cell research. Nat. Rev. Cancer 5, 311–321 (2005). 46. Moghbeli, M., Moghbeli, F., Forghanifard, M. M. & Abbaszadegan, M. R. Cancer
16. Kanwar, S. S., Yu, Y., Nautiyal, J., Patel, B. B. & Majumdar, A. P. The Wnt/beta- stem cell detection and isolation. Med. Oncol. 31, 69 (2014).
catenin pathway regulates growth and maintenance of colonospheres. Mol. 47. Goodell, M. A., Brose, K., Paradis, G., Conner, A. S. & Mulligan, R. C. Isolation and
Cancer 9, 212 (2010). functional properties of murine hematopoietic stem cells that are replicating
17. Li, C. et al. Identification of pancreatic cancer stem cells. Cancer Res. 67, in vivo. J. Exp. Med. 183, 1797–1806 (1996).
1030–1037 (2007). 48. Pattabiraman, D. R. & Weinberg, R. A. Tackling the cancer stem cells—what
18. Wang, J. et al. Notch promotes radioresistance of glioma stem cells. Stem Cells challenges do they pose? Nat. Rev. Drug Discov. 13, 497–512 (2014).
28, 17–28 (2010). 49. Zhou, S. et al. The ABC transporter Bcrp1/ABCG2 is expressed in a wide variety
19. Pardal, R., Clarke, M. F. & Morrison, S. J. Applying the principles of stem-cell of stem cells and is a molecular determinant of the side-population phenotype.
biology to cancer. Nat. Rev. Cancer 3, 895–902 (2003). Nat. Med. 7, 1028–1034 (2001).
20. Hemmati, H. D. et al. Cancerous stem cells can arise from pediatric brain tumors. 50. Moserle, L., Ghisi, M., Amadori, A. & Indraccolo, S. Side population and cancer
Proc. Natl Acad. Sci. USA 100, 15178–15183 (2003). stem cells: therapeutic implications. Cancer Lett. 288, 1–9 (2010).
21. Jin, X., Jin, X. & Kim, H. Cancer stem cells and differentiation therapy. Tumour 51. Haraguchi, N. et al. Characterization of a side population of cancer cells from
Biol. 39, 1010428317729933 (2017). human gastrointestinal system. Stem Cells 24, 506–513 (2006).
22. Huang, Z., Wu, T., Liu, A. Y. & Ouyang, G. Differentiation and transdifferentiation 52. Szotek, P. P. et al. Ovarian cancer side population defines cells with stem cell-like
potentials of cancer stem cells. Oncotarget 6, 39550–39563 (2015). characteristics and Mullerian Inhibiting Substance responsiveness. Proc. Natl
23. Bussolati, B., Bruno, S., Grange, C., Ferrando, U. & Camussi, G. Identification of a Acad. Sci. USA 103, 11154–11159 (2006).
tumor-initiating stem cell population in human renal carcinomas. FASEB J. 22, 53. Feuring-Buske, M. & Hogge, D. E. Hoechst 33342 efflux identifies a subpopula-
3696–3705 (2008). tion of cytogenetically normal CD34(+)CD38(−) progmetabolism, cell growth,
24. Ricci-Vitiani, L. et al. Tumour vascularization via endothelial differentiation of and tumorigenesisenitor cells from patients with acute myeloid leukemia. Blood
glioblastoma stem-like cells. Nature 468, 824–828 (2010). 97, 3882–3889 (2001).
25. Xiong, Y. Q. et al. Human hepatocellular carcinoma tumor-derived endothelial 54. Wang, M., Wang, Y. & Zhong, J. Side population cells and drug resistance in
cells manifest increased angiogenesis capability and drug resistance compared breast cancer. Mol. Med. Rep. 11, 4297–4302 (2015).
with normal endothelial cells. Clin. Cancer Res. 15, 4838–4846 (2009). 55. Ho, M. M., Ng, A. V., Lam, S. & Hung, J. Y. Side population in human lung cancer
26. Chaffer, C. L. & Weinberg, R. A. A perspective on cancer cell metastasis. Science cell lines and tumors is enriched with stem-like cancer cells. Cancer Res. 67,
331, 1559–1564 (2011). 4827–4833 (2007).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
23
56. Wang, J., Guo, L. P., Chen, L. Z., Zeng, Y. X. & Lu, S. H. Identification of cancer 88. Lin, T., Ding, Y. Q. & Li, J. M. Overexpression of Nanog protein is associated
stem cell-like side population cells in human nasopharyngeal carcinoma cell with poor prognosis in gastric adenocarcinoma. Med. Oncol. 29, 878–885
line. Cancer Res. 67, 3716–3724 (2007). (2012).
57. Chiba, T. et al. Side population purified from hepatocellular carcinoma cells 89. Yu, C. C. et al. MicroRNA let-7a represses chemoresistance and tumourigenicity
harbors cancer stem cell-like properties. Hepatology 44, 240–251 (2006). in head and neck cancer via stem-like properties ablation. Oral Oncol. 47,
58. Montanaro, F. et al. Demystifying SP cell purification: viability, yield, and phe- 202–210 (2011).
notype are defined by isolation parameters. Exp. Cell Res. 298, 144–154 (2004). 90. Chiou, S. H. et al. Coexpression of Oct4 and Nanog enhances malignancy in lung
59. Fang, D. et al. A tumorigenic subpopulation with stem cell properties in mela- adenocarcinoma by inducing cancer stem cell-like properties and
nomas. Cancer Res. 65, 9328–9337 (2005). epithelial–mesenchymal transdifferentiation. Cancer Res. 70, 10433–10444
60. Lobo, N. A., Shimono, Y., Qian, D. & Clarke, M. F. The biology of cancer stem cells. (2010).
Annu. Rev. Cell Dev. Biol. 23, 675–699 (2007). 91. Meng, H. M. et al. Over-expression of Nanog predicts tumor progression and
61. Taylor, M. D. et al. Radial glia cells are candidate stem cells of ependymoma. poor prognosis in colorectal cancer. Cancer Biol. Ther. 9, 295–302 (2010).
Cancer Cell 8, 323–335 (2005). 92. Ibrahim, E. E. et al. Embryonic NANOG activity defines colorectal cancer stem
62. O'Brien, C. A., Kreso, A. & Jamieson, C. H. Cancer stem cells and self-renewal. Clin. cells and modulates through AP1- and TCF-dependent mechanisms. Stem Cells
Cancer Res. 16, 3113–3120 (2010). 30, 2076–2087 (2012).
63. van Stijn, A. et al. Differences between the CD34+ and CD34− blast compart- 93. Wang, X. Q. et al. Epigenetic regulation of pluripotent genes mediates stem cell
ments in apoptosis resistance in acute myeloid leukemia. Haematologica 88, features in human hepatocellular carcinoma and cancer cell lines. PLoS ONE 8,
497–508 (2003). e72435 (2013).
64. Chiodi, I., Belgiovine, C., Dona, F., Scovassi, A. I. & Mondello, C. Drug treatment of 94. Jeter, C. R. et al. Functional evidence that the self-renewal gene NANOG reg-
cancer cell lines: a way to select for cancer stem cells? Cancers 3, 1111–1128 ulates human tumor development. Stem Cells 27, 993–1005 (2009).
(2011). 95. Lu, Y. et al. Knockdown of Oct4 and Nanog expression inhibits the stemness of
65. Bhaskara, V. K., Mohanam, I., Rao, J. S. & Mohanam, S. Intermittent hypoxia pancreatic cancer cells. Cancer Lett. 340, 113–123 (2013).
regulates stem-like characteristics and differentiation of neuroblastoma cells. 96. Flandez, M., Guilmeau, S., Blache, P. & Augenlicht, L. H. KLF4 regulation in
PLoS ONE 7, e30905 (2012). intestinal epithelial cell maturation. Exp. Cell Res. 314, 3712–3723 (2008).
66. Liu, W. H. et al. Efficient enrichment of hepatic cancer stem-like cells from a 97. Cho, Y. G. et al. Genetic and epigenetic analysis of the KLF4 gene in gastric
primary rat HCC model via a density gradient centrifugation-centered method. cancer. APMIS 115, 802–808 (2007).
PLoS ONE 7, e35720 (2012). 98. Bianchi, F. et al. Lung cancers detected by screening with spiral computed
67. Rahimi, K., Fuchtbauer, A. C., Fathi, F., Mowla, S. J. & Fuchtbauer, E. M. Isolation of tomography have a malignant phenotype when analyzed by cDNA microarray.
cancer stem cells by selection for miR-302 expressing cells. PeerJ 7, e6635 Clin. Cancer Res. 10, 6023–6028 (2004).
(2019). 99. Li, Q. et al. Dysregulated Kruppel-like factor 4 and vitamin D receptor signaling
68. Eun, K., Ham, S. W. & Kim, H. Cancer stem cell heterogeneity: origin and new contribute to progression of hepatocellular carcinoma. Gastroenterology 143,
perspectives on CSC targeting. BMB Rep. 50, 117–125 (2017). 799–810 (2012). e792.
69. Zhang, H. & Wang, Z. Z. Mechanisms that mediate stem cell self-renewal and 100. Yasunaga, J. et al. Identification of aberrantly methylated genes in association
differentiation. J. Cell. Biochem. 103, 709–718 (2008). with adult T-cell leukemia. Cancer Res. 64, 6002–6009 (2004).
70. Takahashi, K. & Yamanaka, S. Induction of pluripotent stem cells from mouse 101. Tang, H. et al. KLF4 is a tumor suppressor in anaplastic meningioma stem-like
embryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676 cells and human meningiomas. J. Mol. Cell. Biol. 9, 315–324 (2017).
(2006). 102. Ohnishi, S. et al. Downregulation and growth inhibitory effect of epithelial-type
71. Maherali, N. et al. Directly reprogrammed fibroblasts show global epigenetic Kruppel-like transcription factor KLF4, but not KLF5, in bladder cancer. Biochem.
remodeling and widespread tissue contribution. Cell Stem Cell 1, 55–70 (2007). Biophys. Res. Commun. 308, 251–256 (2003).
72. Okita, K., Ichisaka, T. & Yamanaka, S. Generation of germline-competent induced 103. Luo, A. et al. Discovery of Ca2+-relevant and differentiation-associated genes
pluripotent stem cells. Nature 448, 313–317 (2007). downregulated in esophageal squamous cell carcinoma using cDNA microarray.
73. Nichols, J. et al. Formation of pluripotent stem cells in the mammalian embryo Oncogene 23, 1291–1299 (2004).
depends on the POU transcription factor Oct4. Cell 95, 379–391 (1998). 104. Piestun, D. et al. Nanog transforms NIH3T3 cells and targets cell-type restricted
74. Jerabek, S., Merino, F., Scholer, H. R. & Cojocaru, V. OCT4: dynamic DNA binding genes. Biochem. Biophys. Res. Commun. 343, 279–285 (2006).
pioneers stem cell pluripotency. Biochim. Biophys. Acta 1839, 138–154 (2014). 105. Foster, K. W. et al. Oncogene expression cloning by retroviral transduction of
75. Du, Z. et al. Oct4 is expressed in human gliomas and promotes colony formation adenovirus E1A-immortalized rat kidney RK3E cells: transformation of a host
in glioma cells. Glia 57, 724–733 (2009). with epithelial features by c-MYC and the zinc finger protein GKLF. Cell Growth
76. Murakami, S. et al. SRY and OCT4 are required for the acquisition of cancer stem Differ. 10, 423–434 (1999).
cell-like properties and are potential differentiation therapy targets. Stem Cells 106. Riverso, M., Montagnani, V. & Stecca, B. KLF4 is regulated by RAS/RAF/MEK/ERK
33, 2652–2663 (2015). signaling through E2F1 and promotes melanoma cell growth. Oncogene 36,
77. Ponti, D. et al. Isolation and in vitro propagation of tumorigenic breast cancer 3322–3333 (2017).
cells with stem/progenitor cell properties. Cancer Res. 65, 5506–5511 (2005). 107. Tien, Y. T. et al. Downregulation of the KLF4 transcription factor inhibits the
78. Chen, Y. C. et al. Oct-4 expression maintained cancer stem-like properties in proliferation and migration of canine mammary tumor cells. Vet. J. 205, 244–253
lung cancer-derived CD133-positive cells. PLoS ONE 3, e2637 (2008). (2015).
79. Gwak, J. M., Kim, M., Kim, H. J., Jang, M. H. & Park, S. Y. Expression of embryonal 108. Bretones, G., Delgado, M. D. & Leon, J. Myc and cell cycle control. Biochim.
stem cell transcription factors in breast cancer: Oct4 as an indicator for poor Biophys. Acta 1849, 506–516 (2015).
clinical outcome and tamoxifen resistance. Oncotarget 8, 36305–36318 (2017). 109. Dang, C. V. MYC, metabolism, cell growth, and tumorigenesis. Cold Spring Harbor
80. Song, B. et al. OCT4 directly regulates stemness and extracellular matrix-related Perspect. Med. 3, a014217 (2013).
genes in human germ cell tumours. Biochem. Biophys. Res. Commun. 503, 110. Conzen, S. D. et al. Induction of cell cycle progression and acceleration of
1980–1986 (2018). apoptosis are two separable functions of c-Myc: transrepression correlates with
81. Rodriguez-Pinilla, S. M. et al. Sox2: a possible driver of the basal-like phenotype acceleration of apoptosis. Mol. Cell. Biol. 20, 6008–6018 (2000).
in sporadic breast cancer. Mod. Pathol. 20, 474–481 (2007). 111. Galardi, S. et al. Resetting cancer stem cell regulatory nodes upon MYC inhibi-
82. Hagerstrand, D. et al. Identification of a SOX2-dependent subset of tumor- and tion. EMBO Rep. 17, 1872–1889 (2016).
sphere-forming glioblastoma cells with a distinct tyrosine kinase inhibitor sen- 112. Beer, S. et al. Developmental context determines latency of MYC-induced
sitivity profile. Neuro-Oncology 13, 1178–1191 (2011). tumorigenesis. PLoS Biol. 2, e332 (2004).
83. Gangemi, R. M. et al. SOX2 silencing in glioblastoma tumor-initiating cells causes 113. Jacobs, J. J. et al. Bmi-1 collaborates with c-Myc in tumorigenesis by inhibiting c-
stop of proliferation and loss of tumorigenicity. Stem Cells 27, 40–48 (2009). Myc-induced apoptosis via INK4a/ARF. Genes Dev. 13, 2678–2690 (1999).
84. Basu-Roy, U. et al. Sox2 maintains self renewal of tumor-initiating cells in 114. Cartwright, P. et al. LIF/STAT3 controls ES cell self-renewal and pluripotency by a
osteosarcomas. Oncogene 31, 2270–2282 (2012). Myc-dependent mechanism. Development 132, 885–896 (2005).
85. Boumahdi, S. et al. SOX2 controls tumour initiation and cancer stem-cell func- 115. Liu, Z. et al. SOD2 is a C-myc target gene that promotes the migration and
tions in squamous-cell carcinoma. Nature 511, 246–250 (2014). invasion of tongue squamous cell carcinoma involving cancer stem-like cells.
86. Chambers, I. et al. Functional expression cloning of Nanog, a pluripotency Int. J. Biochem. Cell Biol. 60, 139–146 (2015).
sustaining factor in embryonic stem cells. Cell 113, 643–655 (2003). 116. Pfister, S. et al. Outcome prediction in pediatric medulloblastoma based on DNA
87. Nagata, T. et al. Prognostic significance of NANOG and KLF4 for breast cancer. copy-number aberrations of chromosomes 6q and 17q and the MYC and MYCN
Breast Cancer 21, 96–101 (2014). loci. J. Clin. Oncol. 27, 1627–1636 (2009).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
24
117. Rickman, D. S., Schulte, J. H. & Eilers, M. The expanding world of N-MYC-driven 151. Ji, C. et al. Capillary morphogenesis gene 2 maintains gastric cancer stem-like
tumors. Cancer Discov. 8, 150–163 (2018). cell phenotype by activating a Wnt/beta-catenin pathway. Oncogene 37,
118. Hirvonen, H., Hukkanen, V., Salmi, T. T., Pelliniemi, T. T. & Alitalo, R. L-myc and N- 3953–3966 (2018).
myc in hematopoietic malignancies. Leuk. Lymphoma 11, 197–205 (1993). 152. Wang, T. et al. SMYD3 controls a Wnt-responsive epigenetic switch for ASCL2
119. Shachaf, C. M. et al. MYC inactivation uncovers pluripotent differentiation and activation and cancer stem cell maintenance. Cancer Lett. 430, 11–24 (2018).
tumour dormancy in hepatocellular cancer. Nature 431, 1112–1117 (2004). 153. Wang, Y. et al. The long noncoding RNA lncTCF7 promotes self-renewal of
120. Ellwood-Yen, K. et al. Myc-driven murine prostate cancer shares molecular human liver cancer stem cells through activation of Wnt signaling. Cell Stem Cell
features with human prostate tumors. Cancer Cell 4, 223–238 (2003). 16, 413–425 (2015).
121. Logan, C. Y. & Nusse, R. The Wnt signaling pathway in development and disease. 154. Zhu, P. et al. lnc-beta-Catm elicits EZH2-dependent beta-catenin stabilization
Annu. Rev. Cell Dev. Biol. 20, 781–810 (2004). and sustains liver CSC self-renewal. Nat. Struct. Mol. Biol. 23, 631–639 (2016).
122. Kahn, M. Can we safely target the WNT pathway? Nat. Rev. Drug Discov. 13, 155. Chen, Z., Yao, L., Liu, Y. & Zhu, P. LncTIC1 interacts with beta-catenin to drive
513–532 (2014). liver TIC self-renewal and liver tumorigenesis. Cancer Lett. 430, 88–96 (2018).
123. Katoh, M. Canonical and non-canonical WNT signaling in cancer stem cells and 156. Chai, S. et al. Octamer 4/microRNA-1246 signaling axis drives Wnt/beta-catenin
their niches: cellular heterogeneity, omics reprogramming, targeted therapy activation in liver cancer stem cells. Hepatology 64, 2062–2076 (2016).
and tumor plasticity (Review). Int. J. Oncol. 51, 1357–1369 (2017). 157. Mo, X. M., Li, H. H., Liu, M. & Li, Y. T. Downregulation of GSK3beta by miR-544a to
124. Latres, E., Chiaur, D. S. & Pagano, M. The human F box protein beta-Trcp maintain self-renewal ability of lung caner stem cells. Oncol. Lett. 8, 1731–1734
associates with the Cul1/Skp1 complex and regulates the stability of beta- (2014).
catenin. Oncogene 18, 849–854 (1999). 158. Wu, K., Ma, L. & Zhu, J. miR4835p promotes growth, invasion and selfrenewal of
125. Metcalfe, C., Mendoza-Topaz, C., Mieszczanek, J. & Bienz, M. Stability elements in gastric cancer stem cells by Wnt/betacatenin signaling. Mol. Med. Rep. 14,
the LRP6 cytoplasmic tail confer efficient signalling upon DIX-dependent 3421–3428 (2016).
polymerization. J. Cell Sci. 123, 1588–1599 (2010). 159. Ordonez-Moran, P., Dafflon, C., Imajo, M., Nishida, E. & Huelsken, J. HOXA5
126. Tree, D. R. et al. Prickle mediates feedback amplification to generate asymmetric counteracts stem cell traits by inhibiting Wnt signaling in colorectal cancer.
planar cell polarity signaling. Cell 109, 371–381 (2002). Cancer Cell 28, 815–829 (2015).
127. Habas, R., Kato, Y. & He, X. Wnt/Frizzled activation of Rho regulates vertebrate 160. Cai, W. et al. PMP22 regulates self-renewal and chemoresistance of gastric
gastrulation and requires a novel Formin homology protein Daam1. Cell 107, cancer cells. Mol. Cancer Ther. 16, 1187–1198 (2017).
843–854 (2001). 161. Lettini, G. et al. TRAP1 regulates stemness through Wnt/beta-catenin pathway in
128. Kikuchi, A., Yamamoto, H., Sato, A. & Matsumoto, S. New insights into the human colorectal carcinoma. Cell Death Differ. 23, 1792–1803 (2016).
mechanism of Wnt signaling pathway activation. Int. Rev. Cell. Mol. Biol. 291, 162. Lv, Z. et al. Expression and functional regulation of stemness gene Lgr5 in
21–71 (2011). esophageal squamous cell carcinoma. Oncotarget 8, 26492–26504 (2017).
129. Gao, C. & Chen, Y. G. Dishevelled: the hub of Wnt signaling. Cell. Signal. 22, 163. Mu, J. et al. Dickkopf-related protein 2 induces G0/G1 arrest and apoptosis
717–727 (2010). through suppressing Wnt/β-catenin signaling and is frequently methylated in
130. Thompson, J. J. & Williams, C. S. Protein phosphatase 2A in the regulation of Wnt breast cancer. Oncotarget 8, 39443–39459 (2017).
signaling stem cells, and cancer. Genes 9, 121 (2018). 164. Yin, X. et al. DACT1, an antagonist to Wnt/beta-catenin signaling, suppresses
131. Abd El-Rehim, D. & Ali, M. M. Aberrant expression of beta-catenin in invasive tumor cell growth and is frequently silenced in breast cancer. Breast Cancer Res.
ductal breast carcinomas. J. Egypt. Natl Cancer Inst. 21, 185–195 (2009). 15, R23 (2013).
132. Adachi, S. et al. Role of a BCL9-related beta-catenin-binding protein, B9L, in 165. Li, L. et al. The human cadherin 11 is a pro-apoptotic tumor suppressor mod-
tumorigenesis induced by aberrant activation of Wnt signaling. Cancer Res. 64, ulating cell stemness through Wnt/beta-catenin signaling and silenced in
8496–8501 (2004). common carcinomas. Oncogene 31, 3901–3912 (2012).
133. Ishigaki, K. et al. Aberrant localization of beta-catenin correlates with over- 166. Zhu, J. et al. Wnt/beta-catenin pathway mediates (−)-Epigallocatechin-3-gallate
expression of its target gene in human papillary thyroid cancer. J. Clin. Endo- (EGCG) inhibition of lung cancer stem cells. Biochem. Biophys. Res. Commun.
crinol. Metab. 87, 3433–3440 (2002). 482, 15–21 (2017).
134. Kudo, J. et al. Aberrant nuclear localization of beta-catenin without genetic 167. Kim, J. Y. et al. CWP232228 targets liver cancer stem cells through Wnt/beta-
alterations in beta-catenin or Axin genes in esophageal cancer. World J. Surg. catenin signaling: a novel therapeutic approach for liver cancer treatment.
Oncol. 5, 21 (2007). Oncotarget 7, 20395–20409 (2016).
135. Pan, T., Xu, J. & Zhu, Y. Self-renewal molecular mechanisms of colorectal cancer 168. Shi, L., Fei, X., Wang, Z. & You, Y. PI3K inhibitor combined with miR-125b inhi-
stem cells. Int. J. Mol. Med. 39, 9–20 (2017). bitor sensitize TMZ-induced anti-glioma stem cancer effects through inactiva-
136. Mazzoni, S. M. & Fearon, E. R. AXIN1 and AXIN2 variants in gastrointestinal tion of Wnt/beta-catenin signaling pathway. Vitr. Cell Dev. Biol. Anim. 51,
cancers. Cancer Lett. 355, 1–8 (2014). 1047–1055 (2015).
137. Morin, P. J. Beta-catenin signaling and cancer. BioEssays 21, 1021–1030 (1999). 169. DiMeo, T. A. et al. A novel lung metastasis signature links Wnt signaling with
138. Laurent-Puig, P. et al. Genetic alterations associated with hepatocellular carci- cancer cell self-renewal and epithelial-mesenchymal transition in basal-like
nomas define distinct pathways of hepatocarcinogenesis. Gastroenterology 120, breast cancer. Cancer Res. 69, 5364–5373 (2009).
1763–1773 (2001). 170. Li, Q. et al. FZD8, a target of p53, promotes bone metastasis in prostate cancer
139. Clements, W. M. et al. Beta-catenin mutation is a frequent cause of Wnt pathway by activating canonical Wnt/beta-catenin signaling. Cancer Lett. 402, 166–176
activation in gastric cancer. Cancer Res. 62, 3503–3506 (2002). (2017).
140. Liang, J. et al. Mitochondrial PKM2 regulates oxidative stress-induced apoptosis 171. de Lau, W. et al. Lgr5 homologues associate with Wnt receptors and mediate R-
by stabilizing Bcl2. Cell Res. 27, 329–351 (2017). spondin signalling. Nature 476, 293–297 (2011).
141. Zhang, K. et al. WNT/beta-catenin directs self-renewal symmetric cell division of 172. Carmon, K. S., Gong, X., Lin, Q., Thomas, A. & Liu, Q. R-spondins function as
hTERT(high) prostate cancer stem cells. Cancer Res. 77, 2534–2547 (2017). ligands of the orphan receptors LGR4 and LGR5 to regulate Wnt/beta-catenin
142. Hwang, W. L. et al. MicroRNA-146a directs the symmetric division of Snail- signaling. Proc. Natl Acad. Sci. USA 108, 11452–11457 (2011).
dominant colorectal cancer stem cells. Nat. Cell Biol. 16, 268–280 (2014). 173. Todaro, M. et al. CD44v6 is a marker of constitutive and reprogrammed cancer
143. Fang, L. et al. Aberrantly expressed miR-582-3p maintains lung cancer stem cell- stem cells driving colon cancer metastasis. Cell Stem Cell 14, 342–356 (2014).
like traits by activating Wnt/β-catenin signalling. Nat. Commun. 6, 8640 (2015). 174. Matzke-Ogi, A. et al. Inhibition of tumor growth and metastasis in pancreatic
144. Clevers, H. & Nusse, R. Wnt/β-catenin signaling and disease. Cell 149 (2012). cancer models by interference with CD44v6 signaling. Gastroenterology 150,
145. Weinberg, R. A. The many faces of tumor dormancy. APMIS 116, 548–551 (2008). 513–525 (2016). e510.
146. Reya, T. λ et al. Stem cells, cancer, and cancer stem cells. Nature 414, 105 (2001). 175. Su, J., Wu, S., Wu, H., Li, L. & Guo, T. CD44 is functionally crucial for driving lung
147. Giancotti, F. G. Mechanisms governing metastatic dormancy and reactivation. cancer stem cells metastasis through Wnt/beta-catenin-FoxM1-Twist signaling.
Cell 155, 750–764 (2013). Mol. Carcinogen. 55, 1962–1973 (2016).
148. O"Connell, J. T. et al. VEGF-A and Tenascin-C produced by S100A4+ stromal cells 176. Imajo, M., Miyatake, K., Iimura, A., Miyamoto, A. & Nishida, E. A molecular
are important for metastatic colonization. Proc. Natl Acad. Sci. USA 108, mechanism that links Hippo signalling to the inhibition of Wnt/beta-catenin
16002–16007 (2011). signalling. EMBO J. 31, 1109–1122 (2012).
149. Zhao, Z. et al. PKM2 promotes stemness of breast cancer cell by through Wnt/ 177. Diamantopoulou, Z. et al. TIAM1 antagonizes TAZ/YAP both in the destruction
beta-catenin pathway. Tumour Biol. 37, 4223–4234 (2016). complex in the cytoplasm and in the nucleus to inhibit invasion of intestinal
150. Chen, J. F. et al. EZH2 promotes colorectal cancer stem-like cell expansion by epithelial cells. Cancer Cell 31, 621–634 (2017). e626.
activating p21cip1-Wnt/β-catenin signaling. Oncotarget 7, 41540–41558 178. Bray, S. J. Notch signalling: a simple pathway becomes complex. Nat. Rev. Mol.
(2016). Cell. Biol. 7, 678–689 (2006).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
25
179. Schroeter, E. H., Kisslinger, J. A. & Kopan, R. Notch-1 signalling requires ligand- 208. Xie, M. et al. Activation of Notch-1 enhances epithelial–mesenchymal transition
induced proteolytic release of intracellular domain. Nature 393, 382–386 in gefitinib-acquired resistant lung cancer cells. J. Cell. Biochem. 113, 1501–1513
(1998). (2012).
180. Kovall, R. A. More complicated than it looks: assembly of Notch pathway tran- 209. Qiang, L. et al. HIF-1alpha is critical for hypoxia-mediated maintenance of
scription complexes. Oncogene 27, 5099–5109 (2008). glioblastoma stem cells by activating Notch signaling pathway. Cell Death Differ.
181. Ranganathan, P., Weaver, K. L. & Capobianco, A. J. Notch signalling in solid 19, 284–294 (2012).
tumours: a little bit of everything but not all the time. Nat. Rev. Cancer 11, 210. Nwaeburu, C. C., Abukiwan, A., Zhao, Z. & Herr, I. Quercetin-induced miR-200b-
338–351 (2011). 3p regulates the mode of self-renewing divisions in pancreatic cancer. Mol.
182. Zhou, W. et al. The AKT1/NF-kappaB/Notch1/PTEN axis has an important role in Cancer 16, 23 (2017).
chemoresistance of gastric cancer cells. Cell Death Dis. 4, e847 (2013). 211. Lu, J. et al. MiR-26a inhibits stem cell-like phenotype and tumor growth of
183. Wu, F., Stutzman, A. & Mo, Y. Y. Notch signaling and its role in breast cancer. osteosarcoma by targeting Jagged1. Oncogene 36, 231–241 (2017).
Front. Biosci. 12, 4370–4383 (2007). 212. Merchant, A. A. & Matsui, W. Targeting Hedgehog—a cancer stem cell pathway.
184. Zhang, Y., Li, B., Ji, Z. Z. & Zheng, P. S. Notch1 regulates the growth of human Clin. Cancer Res. 16, 3130–3140 (2010).
colon cancers. Cancer 116, 5207–5218 (2010). 213. Binder, M. et al. Functionally distinctive Ptch receptors establish multimodal
185. Gupta, A. et al. Neuroendocrine differentiation in the 12T-10 transgenic prostate hedgehog signaling in the tooth epithelial stem cell niche. Stem Cells 37,
mouse model mimics endocrine differentiation of pancreatic beta cells. Prostate 1238–1248 (2019).
68, 50–60 (2008). 214. Hui, C. C. & Angers, S. Gli proteins in development and disease. Annu. Rev. Cell
186. Lefort, K. et al. Notch1 is a p53 target gene involved in human keratinocyte Dev. Biol. 27, 513–537 (2011).
tumor suppression through negative regulation of ROCK1/2 and MRCKalpha 215. Wang, B., Fallon, J. F. & Beachy, P. A. Hedgehog-regulated processing of Gli3
kinases. Genes Dev. 21, 562–577 (2007). produces an anterior/posterior repressor gradient in the developing vertebrate
187. Konishi, J. et al. Notch3 cooperates with the EGFR pathway to modulate limb. Cell 100, 423–434 (2000).
apoptosis through the induction of bim. Oncogene 29, 589–596 (2010). 216. Sasaki, H., Nishizaki, Y., Hui, C., Nakafuku, M. & Kondoh, H. Regulation of Gli2 and
188. Viatour, P. et al. Notch signaling inhibits hepatocellular carcinoma following Gli3 activities by an amino-terminal repression domain: implication of Gli2 and
inactivation of the RB pathway. J. Exp. Med. 208, 1963–1976 (2011). Gli3 as primary mediators of Shh signaling. Development 126, 3915–3924 (1999).
189. Parr, C., Watkins, G. & Jiang, W. G. The possible correlation of Notch-1 and 217. Li, S. H., Fu, J., Watkins, D. N., Srivastava, R. K. & Shankar, S. Sulforaphane reg-
Notch-2 with clinical outcome and tumour clinicopathological parameters in ulates self-renewal of pancreatic cancer stem cells through the modulation of
human breast cancer. Int. J. Mol. Med. 14, 779–786 (2004). Sonic hedgehog-GLI pathway. Mol. Cell. Biochem. 373, 217–227 (2013).
190. Li, L. et al. Notch signaling pathway networks in cancer metastasis: a new target 218. Petrova, R. & Joyner, A. L. Roles for Hedgehog signaling in adult organ home-
for cancer therapy. Med. Oncol. 34, 180 (2017). ostasis and repair. Development 141, 3445–3457 (2014).
191. Espinoza, I. & Miele, L. Notch inhibitors for cancer treatment. Pharmacol. Ther. 219. Chaudhry, P., Singh, M., Triche, T. J., Guzman, M. & Merchant, A. A. GLI3 repressor
139, 95–110 (2013). determines Hedgehog pathway activation and is required for response to SMO
192. Stylianou, S., Clarke, R. B. & Brennan, K. Aberrant activation of notch signaling in antagonist glasdegib in AML. Blood 129, 3465–3475 (2017).
human breast cancer. Cancer Res. 66, 1517–1525 (2006). 220. Zhou, Q. & Kalderon, D. Hedgehog activates fused through phosphorylation to
193. Harrison, H. et al. Regulation of breast cancer stem cell activity by signaling elicit a full spectrum of pathway responses. Dev. Cell 20, 802–814 (2011).
through the Notch4 receptor. Cancer Res. 70, 709–718 (2010). 221. Robbins, D. J., Fei, D. L. & Riobo, N. A. The Hedgehog signal transduction net-
194. Kang, L. et al. MicroRNA-34a suppresses the breast cancer stem cell-like work. Sci. Signal. 5, re6 (2012).
characteristics by downregulating Notch1 pathway. Cancer Sci. 106, 700–708 222. McMillan, R. & Matsui, W. Molecular pathways: the hedgehog signaling pathway
(2015). in cancer. Clin. Cancer Res. 18, 4883–4888 (2012).
195. Miao, Z. F. et al. DLL4 overexpression increases gastric cancer stem/progenitor 223. Villegas, V. E. et al. Tamoxifen treatment of breast cancer cells: impact on
cell self-renewal ability and correlates with poor clinical outcome via Notch-1 hedgehog/GLI1 signaling. Int. J. Mol. Sci. 17, 308 (2016).
signaling pathway activation. Cancer Med. 6, 245–257 (2017). 224. Pietanza, M. C. et al. A phase I trial of the Hedgehog inhibitor, sonidegib
196. Zhang, C. et al. Actin cytoskeleton regulator Arp2/3 complex is required for DLL1 (LDE225), in combination with etoposide and cisplatin for the initial treatment
activating Notch1 signaling to maintain the stem cell phenotype of glioma of extensive stage small cell lung cancer. Lung Cancer 99, 23–30 (2016).
initiating cells. Oncotarget 8, 33353–33364 (2017). 225. Amantini, C. et al. Capsaicin triggers autophagic cell survival which drives epi-
197. Garcia-Heredia, J. M., Lucena-Cacace, A., Verdugo-Sivianes, E. M., Perez, M. & thelial mesenchymal transition and chemoresistance in bladder cancer cells in
Carnero, A. The Cargo Protein MAP17 (PDZK1IP1) regulates the cancer stem cell an Hedgehog-dependent manner. Oncotarget 7, 50180–50194 (2016).
pool activating the notch pathway by abducting NUMB. Clin. Cancer Res. 23, 226. Xu, Y., An, Y., Wang, X., Zha, W. & Li, X. Inhibition of the Hedgehog pathway
3871–3883 (2017). induces autophagy in pancreatic ductal adenocarcinoma cells. Oncol. Rep. 31,
198. Wang, R. et al. iNOS promotes CD24(+)CD133(+) liver cancer stem cell phe- 707–712 (2014).
notype through a TACE/ADAM17-dependent Notch signaling pathway. Proc. 227. Campbell, V. T. et al. Hedgehog pathway inhibition in chondrosarcoma using
Natl Acad. Sci. USA 115, E10127–E10136 (2018). the smoothened inhibitor IPI-926 directly inhibits sarcoma cell growth. Mol.
199. Lee, S. H. et al. TNFα enhances cancer stem cell-like phenotype via Notch-Hes1 Cancer Ther. 13, 1259–1269 (2014).
activation in oral squamous cell carcinoma cells. Biochem. Biophys. Res. Commun. 228. Zibat, A. et al. Activation of the hedgehog pathway confers a poor prognosis in
424, 58–64 (2012). embryonal and fusion gene-negative alveolar rhabdomyosarcoma. Oncogene
200. Zhang, F. et al. Overexpression of PER3 inhibits self-renewal capability and 29, 6323–6330 (2010).
chemoresistance of colorectal cancer stem-like cells via inhibition of notch and 229. Louis, C. U. & Shohet, J. M. Neuroblastoma: molecular pathogenesis and therapy.
beta-catenin signaling. Oncol. Res. 25, 709–719 (2017). Annu. Rev. Med. 66, 49–63 (2015).
201. Yu, F. et al. Kruppel-like factor 4 (KLF4) is required for maintenance of breast 230. Buonamici, S. et al. Interfering with resistance to smoothened antagonists by
cancer stem cells and for cell migration and invasion. Oncogene 30, 2161–2172 inhibition of the PI3K pathway in medulloblastoma. Sci. Transl. Med. 2, 51ra70
(2011). (2010).
202. Choi, S. et al. BMP-4 enhances epithelial–mesenchymal transition and cancer 231. Yoon, C. et al. CD44 expression denotes a subpopulation of gastric cancer cells
stem cell properties of breast cancer cells via Notch signaling. Sci. Rep. 9, 11724 in which Hedgehog signaling promotes chemotherapy resistance. Clin. Cancer
(2019). Res. 20, 3974–3988 (2014).
203. Buckley, N. E. et al. BRCA1 is a key regulator of breast differentiation through 232. Hahn, H. et al. Mutations of the human homolog of Drosophila patched in the
activation of Notch signalling with implications for anti-endocrine treatment of nevoid basal cell carcinoma syndrome. Cell 85, 841–851 (1996).
breast cancers. Nucleic Acids Res. 41, 8601–8614 (2013). 233. Pietsch, T. et al. Medulloblastomas of the desmoplastic variant carry mutations
204. Rodrigues, M. et al. GLI3 knockdown decreases stemness, cell proliferation and of the human homologue of Drosophila patched. Cancer Res. 57, 2085–2088
invasion in oral squamous cell carcinoma. Int. J. Oncol. 53, 2458–2472 (2018). (1997).
205. Kao, S. H., Wu, K. J. & Lee, W. H. Hypoxia, epithelial–mesenchymal transition, and 234. Taylor, M. D. et al. Mutations in SUFU predispose to medulloblastoma. Nat.
TET-mediated epigenetic changes. J. Clin. Med. 5, (2016). Genet. 31, 306–310 (2002).
206. Sahlgren, C., Gustafsson, M. V., Jin, S., Poellinger, L. & Lendahl, U. Notch signaling 235. Wong, A. J. et al. Increased expression of the epidermal growth factor receptor
mediates hypoxia-induced tumor cell migration and invasion. Proc. Natl Acad. gene in malignant gliomas is invariably associated with gene amplification. Proc.
Sci. USA 105, 6392–6397 (2008). Natl Acad. Sci. USA 84, 6899–6903 (1987).
207. Xing, F. et al. Hypoxia-induced Jagged2 promotes breast cancer metastasis and 236. Zeng, C. et al. SPOP suppresses tumorigenesis by regulating Hedgehog/
self-renewal of cancer stem-like cells. Oncogene 30, 4075–4086 (2011). Gli2 signaling pathway in gastric cancer. J. Exp. Clin. Cancer Res. 33, 75 (2014).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
26
237. Yang, L., Xie, G., Fan, Q. & Xie, J. Activation of the hedgehog-signaling pathway 267. Kim, B. R. et al. RUNX3 suppresses metastasis and stemness by inhibiting
in human cancer and the clinical implications. Oncogene 29, 469–481 (2010). Hedgehog signaling in colorectal cancer. Cell Death Differ. 27, 676-694 (2019).
238. Dierks, C. et al. Expansion of Bcr-Abl-positive leukemic stem cells is dependent 268. Wang, F. et al. Hedgehog signaling regulates epithelial–mesenchymal transition
on Hedgehog pathway activation. Cancer Cell 14, 238–249 (2008). in pancreatic cancer stem-like cells. J. Cancer 7, 408–417 (2016).
239. Zhao, C. et al. Hedgehog signalling is essential for maintenance of cancer stem 269. Zinke, J. et al. Beta-catenin-Gli1 interaction regulates proliferation and tumor
cells in myeloid leukaemia. Nature 458, 776–779 (2009). growth in medulloblastoma. Mol. Cancer 14, 17 (2015).
240. Po, A. et al. Noncanonical GLI1 signaling promotes stemness features and 270. Tang, B. et al. MicroRNA-324-5p regulates stemness, pathogenesis and sensi-
in vivo growth in lung adenocarcinoma. Oncogene 36, 4641–4652 (2017). tivity to bortezomib in multiple myeloma cells by targeting hedgehog signaling.
241. Clement, V., Sanchez, P., de Tribolet, N., Radovanovic, I. & Ruiz i Altaba, A. Int. J. Cancer 142, 109–120 (2018).
HEDGEHOG-GLI1 signaling regulates human glioma growth, cancer stem cell 271. Miele, E. et al. β-Arrestin1-mediated acetylation of Gli1 regulates Hedgehog/Gli
self-renewal, and tumorigenicity. Curr. Biol. 17, 165–172 (2007). signaling and modulates self-renewal of SHH medulloblastoma cancer stem
242. Zhou, A. et al. Gli1-induced deubiquitinase USP48 aids glioblastoma tumor- cells. BMC Cancer 17, 488 (2017).
igenesis by stabilizing Gli1. EMBO Rep. 18, 1318–1330 (2017). 272. Du, W. et al. Targeting the SMO oncogene by miR-326 inhibits glioma biological
243. Zhang, S. et al. Inhibition of CK2alpha down-regulates Hedgehog/Gli signaling behaviors and stemness. Neuro-Oncology 17, 243–253 (2015).
leading to a reduction of a stem-like side population in human lung cancer cells. 273. Chandimali, N. et al. MicroRNA-122 negatively associates with peroxiredoxin-II
PLoS ONE 7, e38996 (2012). expression in human gefitinib-resistant lung cancer stem cells. Cancer Gene Ther.
244. Wen, J. et al. WIP1 modulates responsiveness to Sonic Hedgehog signaling in 26(9–10):292–304(2018).
neuronal precursor cells and medulloblastoma. Oncogene 35, 5552–5564 274. Zhang, Q., Lenardo, M. J. & Baltimore, D. 30 Years of NF-κB: a blossoming of
(2016). relevance to human pathobiology. Cell 168, 37–57 (2017).
245. Pandolfi, S. et al. WIP1 phosphatase modulates the Hedgehog signaling by 275. Hoesel, B. & Schmid, J. A. The complexity of NF-κB signaling in inflammation and
enhancing GLI1 function. Oncogene 32, 4737–4747 (2013). cancer. Mol. Cancer 12, 86 (2013).
246. Raducu, M. et al. SCF (Fbxl17) ubiquitylation of Sufu regulates Hedgehog sig- 276. Hayden, M. S. & Ghosh, S. Shared principles in NF-kappaB signaling. Cell 132,
naling and medulloblastoma development. EMBO J. 35, 1400–1416 (2016). 344–362 (2008).
247. Yang, Y. et al. RARalpha2 expression confers myeloma stem cell features. Blood 277. May, M. J. & Ghosh, S. Rel/NF-kappa B and I kappa B proteins: an overview.
122, 1437–1447 (2013). Semin. Cancer Biol. 8, 63–73 (1997).
248. Song, K. B., Liu, W. J. & Jia, S. S. miR-219 inhibits the growth and metastasis of 278. Novack, D. V. Role of NF-κB in the skeleton. Cell Res. 21, 169–182 (2011).
TSCC cells by targeting PRKCI. Int. J. Clin. Exp. Med. 7, 2957–2965 (2014). 279. Perkins, N. D. & Gilmore, T. D. Good cop, bad cop: the different faces of NF-
249. Justilien, V. et al. The PRKCI and SOX2 oncogenes are coamplified and cooperate kappaB. Cell Death Differ. 13, 759–772 (2006).
to activate Hedgehog signaling in lung squamous cell carcinoma. Cancer Cell 25, 280. Sun, S. C. Non-canonical NF-kappaB signaling pathway. Cell Res. 21, 71–85
139–151 (2014). (2011).
250. Airoldi, I. et al. Interleukin-27 re-educates intratumoral myeloid cells and down- 281. Xiao, G., Harhaj, E. W. & Sun, S. C. NF-kappaB-inducing kinase regulates the
regulates stemness genes in non-small cell lung cancer. Oncotarget 6, processing of NF-kappaB2 p100. Mol. Cell 7, 401–409 (2001).
3694–3708 (2015). 282. Greten, F. R. et al. IKKbeta links inflammation and tumorigenesis in a mouse
251. Li, C. et al. GALNT1-mediated glycosylation and activation of sonic hedgehog model of colitis-associated cancer. Cell 118, 285–296 (2004).
signaling maintains the self-renewal and tumor-initiating capacity of bladder 283. Taniguchi, K. & Karin, M. NF-kappaB, inflammation, immunity and cancer:
cancer stem cells. Cancer Res. 76, 1273–1283 (2016). coming of age. Nat. Rev. Immunol. 18, 309–324 (2018).
252. Memmi, E. M. et al. p63 sustains self-renewal of mammary cancer stem cells 284. Denz, U., Haas, P. S., Wasch, R., Einsele, H. & Engelhardt, M. State of the art
through regulation of Sonic Hedgehog signaling. Proc. Natl Acad. Sci. USA 112, therapy in multiple myeloma and future perspectives. Eur. J. Cancer 42,
3499–3504 (2015). 1591–1600 (2006).
253. Han, B. et al. FOXC1 activates smoothened-independent hedgehog signaling in 285. Karin, M. NF-kappaB as a critical link between inflammation and cancer. Cold
basal-like breast cancer. Cell Rep. 13, 1046–1058 (2015). Spring Harb. Perspect. Biol. 1, a000141 (2009).
254. Wu, J. et al. The long non-coding RNA LncHDAC2 drives the self-renewal of liver 286. Prasad, S., Ravindran, J. & Aggarwal, B. B. NF-kappaB and cancer: how intimate is
cancer stem cells via activation of Hedgehog signaling. J. Hepatol. 70, 918–929 this relationship. Mol. Cell. Biochem. 336, 25–37 (2010).
(2019). 287. Brown, M., Cohen, J., Arun, P., Chen, Z. & Van Waes, C. NF-kappaB in carcinoma
255. Valenti, G. et al. Cancer stem cells regulate cancer-associated fibroblasts via therapy and prevention. Expert Opin. Ther. Targets 12, 1109–1122 (2008).
activation of hedgehog signaling in mammary gland tumors. Cancer Res. 77, 288. Hinz, M. et al. Nuclear factor kappaB-dependent gene expression profiling of
2134–2147 (2017). Hodgkin’s disease tumor cells, pathogenetic significance, and link to con-
256. Gu, D. et al. Combining hedgehog signaling inhibition with focal irradiation on stitutive signal transducer and activator of transcription 5a activity. J. Exp. Med.
reduction of pancreatic cancer metastasis. Mol. Cancer Ther. 12, 1038–1048 196, 605–617 (2002).
(2013). 289. Richmond, A. Nf-kappa B chemokine gene transcription and tumour growth.
257. Joost, S. et al. GLI1 inhibition promotes epithelial-to-mesenchymal transition in Nat. Rev. Immunol. 2, 664–674 (2002).
pancreatic cancer cells. Cancer Res. 72, 88–99 (2012). 290. Gonzalez-Torres, C. et al. NF-kappaB participates in the stem cell phenotype of
258. Liu, S. et al. Hedgehog signaling and Bmi-1 regulate self-renewal of normal and ovarian cancer cells. Arch. Med. Res. 48, 343–351 (2017).
malignant human mammary stem cells. Cancer Res. 66, 6063–6071 (2006). 291. Vazquez-Santillan, K. et al. NF-kappaBeta-inducing kinase regulates stem cell
259. Song, Z. et al. Sonic hedgehog pathway is essential for maintenance of cancer phenotype in breast cancer. Sci. Rep. 6, 37340 (2016).
stem-like cells in human gastric cancer. PLoS ONE 6, e17687 (2011). 292. Kong, L. et al. Overexpression of SDF-1 activates the NF-kappaB pathway to
260. Takahashi, T. et al. Cyclopamine induces eosinophilic differentiation and upre- induce epithelial to mesenchymal transition and cancer stem cell-like pheno-
gulates CD44 expression in myeloid leukemia cells. Leuk. Res. 35, 638–645 types of breast cancer cells. Int. J. Oncol. 48, 1085–1094 (2016).
(2011). 293. Wang, D., Fu, L., Sun, H., Guo, L. & DuBois, R. N. Prostaglandin E2 promotes
261. Takebe, N., Harris, P. J., Warren, R. Q. & Ivy, S. P. Targeting cancer stem cells by colorectal cancer stem cell expansion and metastasis in mice. Gastroenterology
inhibiting Wnt, Notch, and Hedgehog pathways. Nat. Rev. Clin. Oncol. 8, 97–106 149, 1884–1895 (2015). e1884.
(2011). 294. Smith, H. A. & Kang, Y. The metastasis-promoting roles of tumor-associated
262. Tanaka, H. et al. The Hedgehog signaling pathway plays an essential role in immune cells. J. Mol. Med. 91, 411–429 (2013).
maintaining the CD44+CD24−/low subpopulation and the side population of 295. Zhang, L. et al. CCL21/CCR7 axis contributed to CD133+ pancreatic cancer
breast cancer cells. Anticancer Res. 29, 2147–2157 (2009). stem-like cell metastasis via EMT and Erk/NF-κB pathway. PLoS ONE 11,
263. Kong, Y. et al. Twist1 and Snail link Hedgehog signaling to tumor-initiating cell- e0158529 (2016).
like properties and acquired chemoresistance independently of ABC transpor- 296. Wang, X. et al. Bmi-1 regulates stem cell-like properties of gastric cancer cells via
ters. Stem Cells 33, 1063–1074 (2015). modulating miRNAs. J. Hematol. Oncol. 9, 90 (2016).
264. Zhu, R. et al. TSPAN8 promotes cancer cell stemness via activation of sonic 297. Li, B. et al. miR-221/222 promote cancer stem-like cell properties and tumor
Hedgehog signaling. Nat. Commun. 10, 2863 (2019). growth of breast cancer via targeting PTEN and sustained Akt/NF-kappaB/COX-2
265. Guo, E., Liu, H. & Liu, X. Overexpression of SCUBE2 inhibits proliferation, activation. Chem. Biol. Interact. 277, 33–42 (2017).
migration, and invasion in glioma cells. Oncol. Res. 25, 437–444 (2017). 298. Karanikas, V. et al. Foxp3 expression in human cancer cells. J. Transl. Med. 6, 19
266. Tiberi, L. et al. A BCL6/BCOR/SIRT1 complex triggers neurogenesis and sup- (2008).
presses medulloblastoma by repressing Sonic Hedgehog signaling. Cancer Cell 299. Liu, S. et al. FOXP3 inhibits cancer stem cell self-renewal via transcriptional
26, 797–812 (2014). repression of COX2 in colorectal cancer cells. Oncotarget 8, 44694–44704 (2017).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
27
300. Yang, X. et al. MiR-491 attenuates cancer stem cells-like properties of hepato- 330. Kim, S. Y. et al. Role of the IL-6-JAK1-STAT3-Oct-4 pathway in the conversion of
cellular carcinoma by inhibition of GIT-1/NF-kappaB-mediated EMT. Tumour Biol. non-stem cancer cells into cancer stem-like cells. Cell. Signal. 25, 961–969 (2013).
37, 201–209 (2016). 331. Ruan, Z., Yang, X. & Cheng, W. OCT4 accelerates tumorigenesis through acti-
301. Han, D. et al. Disulfiram inhibits TGF-β-induced epithelial–mesenchymal tran- vating JAK/STAT signaling in ovarian cancer side population cells. Cancer
sition and stem-like features in breast cancer via ERK/NF-κB/Snail pathway. Manag. Res. 11, 389–399 (2019).
Oncotarget 6, 40907–40919 (2015). 332. Marotta, L. L. et al. The JAK2/STAT3 signaling pathway is required for growth of
302. Burnett, J. P. et al. Sulforaphane enhances the anticancer activity of taxanes CD44(+)CD24(−) stem cell-like breast cancer cells in human tumors. J. Clin.
against triple negative breast cancer by killing cancer stem cells. Cancer Lett. Invest. 121, 2723–2735 (2011).
394, 52–64 (2017). 333. Zhang, X. et al. Human colorectal cancer-derived mesenchymal stem cells
303. Marquardt, J. U. et al. Curcumin effectively inhibits oncogenic NF-kappaB sig- promote colorectal cancer progression through IL-6/JAK2/STAT3 signaling. Cell
naling and restrains stemness features in liver cancer. J. Hepatol. 63, 661–669 Death Dis. 9, 25 (2018).
(2015). 334. Zhou, B. et al. Erythropoietin promotes breast tumorigenesis through tumor-
304. Levy, D. E. & Darnell, J. E. Jr. Stats: transcriptional control and biological impact. initiating cell self-renewal. J. Clin. Invest. 124, 553–563 (2014).
Nat. Rev. Mol. Cell. Biol. 3, 651–662 (2002). 335. Almiron Bonnin, D. A. et al. Secretion-mediated STAT3 activation promotes self-
305. O’Shea, J. J., Gadina, M. & Schreiber, R. D. Cytokine signaling in 2002: new renewal of glioma stem-like cells during hypoxia. Oncogene 37, 1107–1118
surprises in the Jak/Stat pathway. Cell 109(Suppl.), S121–S131 (2002). (2018).
306. Ihle, J. N. & Kerr, I. M. Jaks and Stats in signaling by the cytokine receptor 336. Jia, H. et al. The LIM protein AJUBA promotes colorectal cancer cell survival
superfamily. Trends Genet. 11, 69–74 (1995). through suppression of JAK1/STAT1/IFIT2 network. Oncogene 36, 2655–2666
307. Haan, C., Kreis, S., Margue, C. & Behrmann, I. Jaks and cytokine receptors—an (2017).
intimate relationship. Biochem. Pharmacol. 72, 1538–1546 (2006). 337. Che, S. et al. miR-30 overexpression promotes glioma stem cells by regulating
308. Yoshimura, A. et al. A novel cytokine-inducible gene CIS encodes an SH2- Jak/STAT3 signaling pathway. Tumour Biol. 36, 6805–6811 (2015).
containing protein that binds to tyrosine-phosphorylated interleukin 3 and 338. Yang, Y. et al. MicroRNA-218 functions as a tumor suppressor in lung cancer by
erythropoietin receptors. EMBO J. 14, 2816–2826 (1995). targeting IL-6/STAT3 and negatively correlates with poor prognosis. Mol. Cancer
309. Quintás-Cardama, A. & Verstovsek, S. Molecular pathways: Jak/STAT pathway: 16, 141 (2017).
mutations, inhibitors, and resistance. Clin. Cancer Res. 19, 1933–1940 (2013). 339. Frolik, C. A., Dart, L. L., Meyers, C. A., Smith, D. M. & Sporn, M. B. Purification and
310. Stroud, R. M. & Wells, J. A. Mechanistic diversity of cytokine receptor signaling initial characterization of a type beta transforming growth factor from human
across cell membranes. Sci. STKE 2004, re7 (2004). placenta. Proc. Natl Acad. Sci. USA 80, 3676–3680 (1983).
311. Chikuma, S., Kanamori, M., Mise-Omata, S. & Yoshimura, A. Suppressors of 340. Weiss, A. & Attisano, L. The TGFbeta superfamily signaling pathway. Wiley
cytokine signaling: potential immune checkpoint molecules for cancer immu- Interdiscip. Rev. Dev. Biol. 2, 47–63 (2013).
notherapy. Cancer Sci. 108, 574–580 (2017). 341. Brown, J. A. et al. TGF-beta-induced quiescence mediates chemoresistance of
312. Leonard, W. J. & O'Shea, J. J. Jaks and STATs: biological implications. Annu. Rev. tumor-propagating cells in squamous cell carcinoma. Cell Stem Cell 21, 650–664
Immunol. 16, 293–322 (1998). (2017). e658.
313. Constantinescu, S. N., Girardot, M. & Pecquet, C. Mining for JAK-STAT mutations 342. Moustakas, A., Souchelnytskyi, S. & Heldin, C. H. Smad regulation in TGF-beta
in cancer. Trends Biochem. Sci. 33, 122–131 (2008). signal transduction. J. Cell Sci. 114, 4359–4369 (2001).
314. Kralovics, R. et al. A gain-of-function mutation of JAK2 in myeloproliferative 343. Massague, J., Seoane, J. & Wotton, D. Smad transcription factors. Genes Dev. 19,
disorders. N. Engl. J. Med. 352, 1779–1790 (2005). 2783–2810 (2005).
315. Kiladjian, J. J. The spectrum of JAK2-positive myeloproliferative neoplasms. 344. Romero, D. et al. Dickkopf-3 regulates prostate epithelial cell acinar morpho-
Hematol. Am. Soc. Hematol. Educ. Program 2012, 561–566 (2012). genesis and prostate cancer cell invasion by limiting TGF-beta-dependent
316. Levine, R. L. et al. Activating mutation in the tyrosine kinase JAK2 in poly- activation of matrix metalloproteases. Carcinogenesis 37, 18–29 (2016).
cythemia vera, essential thrombocythemia, and myeloid metaplasia with mye- 345. Kaowinn, S. et al. Cancer upregulated gene 2 (CUG2), a novel oncogene, pro-
lofibrosis. Cancer Cell 7, 387–397 (2005). motes stemness-like properties via the NPM1-TGF-beta signaling axis. Biochem.
317. Roder, S., Steimle, C., Meinhardt, G. & Pahl, H. L. STAT3 is constitutively active in Biophys. Res. Commun. 514, 1278–1284 (2019).
some patients with Polycythemia rubra vera. Exp. Hematol. 29, 694–702 (2001). 346. Xia, W. et al. Smad inhibitor induces CSC differentiation for effective chemo-
318. James, C. et al. A unique clonal JAK2 mutation leading to constitutive signalling sensitization in cyclin D1- and TGF-beta/Smad-regulated liver cancer stem cell-
causes polycythaemia vera. Nature 434, 1144–1148 (2005). like cells. Oncotarget 8, 38811–38824 (2017).
319. Song, J. I. & Grandis, J. R. STAT signaling in head and neck cancer. Oncogene 19, 347. Wang, X. H. et al. TGF-beta1 pathway affects the protein expression of many
2489–2495 (2000). signaling pathways, markers of liver cancer stem cells, cytokeratins, and TERT in
320. van der Zee, M. et al. IL6/JAK1/STAT3 signaling blockade in endometrial cancer liver cancer HepG2 cells. Tumour Biol. 37, 3675–3681 (2016).
affects the ALDHhi/CD126+ stem-like component and reduces tumor burden. 348. Rodriguez-Garcia, A. et al. TGF-beta1 targets Smad, p38 MAPK, and PI3K/Akt
Cancer Res. 75, 3608–3622 (2015). signaling pathways to induce PFKFB3 gene expression and glycolysis in glio-
321. Lam, L. T. et al. Cooperative signaling through the signal transducer and acti- blastoma cells. FEBS J. 284, 3437–3454 (2017).
vator of transcription 3 and nuclear factor-{kappa}B pathways in subtypes of 349. Muthusamy, B. P. et al. ShcA protects against epithelial–mesenchymal transition
diffuse large B-cell lymphoma. Blood 111, 3701–3713 (2008). through compartmentalized inhibition of TGF-beta-induced Smad activation.
322. Jiang, C. et al. miR-500a-3p promotes cancer stem cells properties via STAT3 PLoS Biol. 13, e1002325 (2015).
pathway in human hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 36, 99 350. Jiang, F. et al. The repressive effect of miR-148a on TGF beta-SMADs signal
(2017). pathway is involved in the glabridin-induced inhibition of the cancer stem
323. Kanno, H., Sato, H., Yokoyama, T. A., Yoshizumi, T. & Yamada, S. The VHL tumor cells-like properties in hepatocellular carcinoma cells. PLoS ONE 9, e96698
suppressor protein regulates tumorigenicity of U87-derived glioma stem-like (2014).
cells by inhibiting the JAK/STAT signaling pathway. Int. J. Oncol. 42, 881–886 351. Yu, D., Shin, H. S., Lee, Y. S. & Lee, Y. C. miR-106b modulates cancer stem cell
(2013). characteristics through TGF-beta/Smad signaling in CD44-positive gastric cancer
324. Karunanithi, S. et al. RBP4-STRA6 pathway drives cancer stem cell maintenance cells. Lab. Invest. 94, 1370–1381 (2014).
and mediates high-fat diet-induced colon carcinogenesis. Stem Cell Rep. 9, 352. Tasian, S. K., Teachey, D. T. & Rheingold, S. R. Targeting the PI3K/mTOR pathway
438–450 (2017). in pediatric hematologic malignancies. Front. Oncol. 4, 108 (2014).
325. Yu, H., Pardoll, D. & Jove, R. STATs in cancer inflammation and immunity: a 353. Vanhaesebroeck, B., Guillermet-Guibert, J., Graupera, M. & Bilanges, B. The
leading role for STAT3. Nat. Rev. Cancer 9, 798–809 (2009). emerging mechanisms of isoform-specific PI3K signalling. Nat. Rev. Mol. Cell.
326. Rajala, H. L. et al. Discovery of somatic STAT5b mutations in large granular Biol. 11, 329–341 (2010).
lymphocytic leukemia. Blood 121, 4541–4550 (2013). 354. Wang, Q., Chen, X. & Hay, N. Akt as a target for cancer therapy: more is not
327. Chambers, I. The molecular basis of pluripotency in mouse embryonic stem always better (lessons from studies in mice). Br. J. Cancer 117, 159–163
cells. Cloning Stem Cells 6, 386–391 (2004). (2017).
328. Zhou, J. et al. Activation of the PTEN/mTOR/STAT3 pathway in breast cancer 355. Loewith, R. et al. Two TOR complexes, only one of which is rapamycin sensitive,
stem-like cells is required for viability and maintenance. Proc. Natl Acad. Sci. USA have distinct roles in cell growth control. Mol. Cell 10, 457–468 (2002).
104, 16158–16163 (2007). 356. Kim, D. H. et al. mTOR interacts with raptor to form a nutrient-sensitive complex
329. Yang, L. et al. IL-10 derived from M2 macrophage promotes cancer stemness via that signals to the cell growth machinery. Cell 110, 163–175 (2002).
JAK1/STAT1/NF-kappaB/Notch1 pathway in non-small cell lung cancer. Int. J. 357. Sancak, Y. et al. PRAS40 is an insulin-regulated inhibitor of the mTORC1 protein
cancer 145, 1099–1110 (2019). kinase. Mol. Cell 25, 903–915 (2007).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
28
358. Knowles, M. A., Platt, F. M., Ross, R. L. & Hurst, C. D. Phosphatidylinositol 3-kinase 384. Houseknecht, K. L., Cole, B. M. & Steele, P. J. Peroxisome proliferator-activated
(PI3K) pathway activation in bladder cancer. Cancer metastasis Rev. 28, 305–316 receptor gamma (PPARgamma) and its ligands: a review. Domest. Anim. Endo-
(2009). crinol. 22, 1–23 (2002).
359. Duan, S. et al. PTEN deficiency reprogrammes human neural stem cells towards 385. Yousefnia, S., Momenzadeh, S., Seyed Forootan, F., Ghaedi, K. & Nasr Esfahani, M.
a glioblastoma stem cell-like phenotype. Nat. Commun. 6, 10068 (2015). H. The influence of peroxisome proliferator-activated receptor gamma (PPAR-
360. Yuzugullu, H. et al. A PI3K p110beta-Rac signalling loop mediates Pten-loss- gamma) ligands on cancer cell tumorigenicity. Gene 649, 14–22 (2018).
induced perturbation of haematopoiesis and leukaemogenesis. Nat. Commun. 6, 386. Kramer, K., Wu, J. & Crowe, D. L. Tumor suppressor control of the cancer stem
8501 (2015). cell niche. Oncogene 35, 4165–4178 (2016).
361. Fumarola, C., Bonelli, M. A., Petronini, P. G. & Alfieri, R. R. Targeting PI3K/AKT/ 387. Wang, Y. et al. The combinatory effects of PPAR-gamma agonist and survivin
mTOR pathway in non small cell lung cancer. Biochem. Pharmacol. 90, 197–207 inhibition on the cancer stem-like phenotype and cell proliferation in bladder
(2014). cancer cells. Int. J. Mol. Med. 34, 262–268 (2014).
362. Dey, N., De, P. & Leyland-Jones, B. PI3K-AKT-mTOR inhibitors in breast cancers: 388. Basu-Roy, U. et al. PPARgamma agonists promote differentiation of cancer stem
from tumor cell signaling to clinical trials. Pharmacol. Ther.175, 91–106 (2017). cells by restraining YAP transcriptional activity. Oncotarget 7,60954–60970
363. Offermann, A. et al. MED15 overexpression in prostate cancer arises during (2016).
androgen deprivation therapy via PI3K/mTOR signaling. Oncotarget 8, 389. Giampietri, C. et al. Lipid storage and autophagy in melanoma cancer cells. Int. J.
7964–7976 (2017). Mol. Sci. 18, (2017).
364. Giulino-Roth, L. et al. Inhibition of Hsp90 suppresses PI3K/AKT/mTOR signaling 390. Deng, X. et al. Ovarian cancer stem cells induce the M2 polarization of mac-
and has antitumor activity in Burkitt lymphoma. Mol. cancer therapeutics 16, rophages through the PPARgamma and NF-kappaB pathways. Int. J. Mol. Med.
1779–1790 (2017). 36, 449–454 (2015).
365. Zaidi, A. H. et al. PI3K/mTOR dual inhibitor, LY3023414, demonstrates potent 391. Bigoni-Ordonez, G. D. et al. Molecular iodine inhibits the expression of stemness
antitumor efficacy against esophageal adenocarcinoma in a rat model. Ann. markers on cancer stem-like cells of established cell lines derived from cervical
Surg. 266, 91–98 (2017). cancer. BMC Cancer 18, 928 (2018).
366. Karki, R., Malireddi, R. K. S., Zhu, Q. & Kanneganti, T. D. NLRC3 regulates cellular 392. Liu, L. et al. Inhibition of oxidative stress-elicited AKT activation facilitates
proliferation and apoptosis to attenuate the development of colorectal cancer. PPARgamma agonist-mediated inhibition of stem cell character and tumor
Cell Cycle 16, 1243–1251 (2017). growth of liver cancer cells. PLoS ONE 8, e73038 (2013).
367. Deng, J. et al. Inhibition of PI3K/Akt/mTOR signaling pathway alleviates 393. Rinaldi, L. et al. Loss of Dnmt3a and Dnmt3b does not affect epidermal
ovarian cancer chemoresistance through reversing epithelial-mesenchymal homeostasis but promotes squamous transformation through PPAR-gamma.
transition and decreasing cancer stem cell marker expression. BMC Cancer 19, eLife 6, pii: e21697 (2017).
618 (2019). 394. Wang, H., Zheng, S., Tu, Y. & Zhang, Y. Screening and identification of novel
368. Chang, L. et al. Acquisition of epithelial–mesenchymal transition and cancer drug-resistant genes in CD133+ and CD133− lung adenosarcoma cells using
stem cell phenotypes is associated with activation of the PI3K/Akt/mTOR cDNA microarray. Chin. J. Lung Cancer 17, 437–443 (2014).
pathway in prostate cancer radioresistance. Cell death Dis. 4, e875 (2013). 395. Lebleu, V. S. et al. PGC-1α mediates mitochondrial biogenesis and oxidative
369. Fitzgerald, T. L. et al. Roles of EGFR and KRAS and their downstream signaling phosphorylation in cancer cells to promote metastasis. Nat. Cell Biol. 16,
pathways in pancreatic cancer and pancreatic cancer stem cells. Adv. Biol. Regul. 992–1003 (2014).
59, 65–81 (2015). 396. Wang, X. et al. AMPK promotes SPOP-mediated NANOG degradation to regulate
370. Dubrovska, A. et al. The role of PTEN/Akt/PI3K signaling in the maintenance and prostate cancer cell stemness. Dev. Cell. 48, 345–360 (2019). e7.
viability of prostate cancer stem-like cell populations. Proc. Natl Acad. Sci. USA 397. Kim, J. H. et al. Effects of metformin on colorectal cancer stem cells depend on
106, 268–273 (2009). alterations in glutamine metabolism. Sci. Rep. 8, 409 (2018).
371. Keysar, S. B. et al. Regulation of head and neck squamous cancer stem cells by 398. Maehara, O. et al. Metformin regulates the expression of CD133 through the
PI3K and SOX2. J. Natl Cancer Inst. 109 (2017). AMPK-CEBPβ pathway in hepatocellular carcinoma cell lines. Neoplasia 21,
372. Yang, C. et al. Downregulation of cancer stem cell properties via mTOR signaling 545–556 (2019).
pathway inhibition by rapamycin in nasopharyngeal carcinoma. Int. J. Oncol. 47, 399. Katoh, M. Networking of WNT, FGF, Notch, BMP, and Hedgehog signaling
909–917 (2015). pathways during carcinogenesis. Stem Cell Rev. 3, 30–38 (2007).
373. Huang, E. H. et al. Aldehyde dehydrogenase 1 is a marker for normal and 400. Schon, S. et al. Beta-catenin regulates NF-kappaB activity via TNFRSF19 in col-
malignant human colonic stem cells (SC) and tracks SC overpopulation during orectal cancer cells. Int. J. Cancer 135, 1800–1811 (2014).
colon tumorigenesis. Cancer Res. 69, 3382–3389 (2009). 401. Liu, D. et al. Reduced CD146 expression promotes tumorigenesis and cancer
374. Nishitani, S., Horie, M., Ishizaki, S. & Yano, H. Branched chain amino acid sup- stemness in colorectal cancer through activating Wnt/beta-catenin signaling.
presses hepatocellular cancer stem cells through the activation of mammalian Oncotarget 7, 40704–40718 (2016).
target of rapamycin. PLoS ONE 8, e82346 (2013). 402. Du, Q. et al. Wnt/beta-catenin signaling regulates cytokine-induced human
375. Karki, R. et al. NLRC3 is an inhibitory sensor of PI3K-mTOR pathways in cancer. inducible nitric oxide synthase expression by inhibiting nuclear factor-kappaB
Nature 540, 583–587 (2016). activation in cancer cells. Cancer Res. 69, 3764–3771 (2009).
376. Chen, W. J. & Huang, R. F. S. Low folate stress reprograms cancer stem cell-like 403. Moreau, M., Mourah, S. & Dosquet, C. Beta-catenin and NF-kappaB cooperate to
potentials and bioenergetics metabolism through activation of mTOR signaling regulate the uPA/uPAR system in cancer cells. Int. J. Cancer 128, 1280–1292
pathway to promote in vitro invasion and in vivo tumorigenicity of lung cancers. (2011).
J. Nutr. Biochem. S0955286316306994 (2017). 404. Thyssen, G. et al. LZTS2 is a novel beta-catenin-interacting protein and regulates
377. Bonuccelli, G., Sotgia, F. & Lisanti, M. P. Matcha green tea (MGT) inhibits the the nuclear export of beta-catenin. Mol. Cell. Biol. 26, 8857–8867 (2006).
propagation of cancer stem cells (CSCs), by targeting mitochondrial metabo- 405. Gwak, J. et al. Polysiphonia japonica extract suppresses the Wnt/beta-catenin
lism, glycolysis and multiple cell signalling pathways. Aging (Albany, NY) pathway in colon cancer cells by activation of NF-kappaB. Int. J. Mol. Med. 17,
10:1867–1883 (2018). 1005–1010 (2006).
378. Issemann, I. & Green, S. Activation of a member of the steroid hormone receptor 406. Albanese, C. et al. IKKalpha regulates mitogenic signaling through transcrip-
superfamily by peroxisome proliferators. Nature 347, 645–650 (1990). tional induction of cyclin D1 via Tcf. Mol. Biol. Cell 14, 585–599 (2003).
379. Peters, J. M., Shah, Y. M. & Gonzalez, F. J. The role of peroxisome proliferator- 407. Noubissi, F. K. et al. Wnt signaling stimulates transcriptional outcome of the
activated receptors in carcinogenesis and chemoprevention. Nat. Rev. Cancer Hedgehog pathway by stabilizing GLI1 mRNA. Cancer Res. 69, 8572–8578 (2009).
12, 181–195 (2012). 408. Regan, J. L. et al. Non-canonical hedgehog signaling is a positive regulator of the
380. Kuenzli, S. & Saurat, J. H. Peroxisome proliferator-activated receptors in cuta- WNT pathway and is required for the survival of colon cancer stem cells. Cell
neous biology. Br. J. Dermatol. 149, 229–236 (2003). Rep. 21, 2813–2828 (2017).
381. Elbrecht, A. et al. Molecular cloning, expression and characterization of human 409. Quan, X. X. et al. Targeting Notch1 and IKKα enhanced NF-κB activation in
peroxisome proliferator activated receptors gamma 1 and gamma 2. Biochem. CD133(+) skin cancer stem cells. Mol. Cancer Ther. 17, 2034–2048 (2018).
Biophys. Res. Commun. 224, 431–437 (1996). 410. Liu, R. Y. et al. JAK/STAT3 signaling is required for TGF-beta-induced
382. Tyagi, S., Gupta, P., Saini, A. S., Kaushal, C. & Sharma, S. The peroxisome epithelial-mesenchymal transition in lung cancer cells. Int. J. Oncol. 44,
proliferator-activated receptor: a family of nuclear receptors role in various 1643–1651 (2014).
diseases. J. Adv. Pharm. Technol. Res. 2, 236–240 (2011). 411. Luo, Y. et al. Non-CSCs nourish CSCs through interleukin-17E-mediated activa-
383. Sertznig, P. & Reichrath, J. Peroxisome proliferator-activated receptors (PPARs) in tion of NF-kappaB and JAK/STAT3 signaling in human hepatocellular carcinoma.
dermatology: challenge and promise. Dermatol. Endocrinol. 3, 130–135 (2011). Cancer Lett. 375, 390–399 (2016).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
29
412. Wu, W. et al. LncRNA NKILA suppresses TGF-beta-induced epithelial- 443. Samanta, D., Gilkes, D. M., Chaturvedi, P., Xiang, L. & Semenza, G. L. Hypoxia-
mesenchymal transition by blocking NF-kappaB signaling in breast cancer. Int. inducible factors are required for chemotherapy resistance of breast cancer
J. Cancer 143, 2213–2224 (2018). stem cells. Proc. Natl Acad. Sci. USA 111, E5429–E5438 (2014).
413. Serra, R., Easter, S. L., Jiang, W. & Baxley, S. E. Wnt5a as an effector of TGFbeta in 444. Heddleston, J. M., Li, Z., McLendon, R. E., Hjelmeland, A. B. & Rich, J. N. The
mammary development and cancer. J. Mammary Gland Biol. Neoplasia 16, hypoxic microenvironment maintains glioblastoma stem cells and promotes
157–167 (2011). reprogramming towards a cancer stem cell phenotype. Cell Cycle 8, 3274–3284
414. Zuo, M. et al. Novel therapeutic strategy targeting the Hedgehog signalling (2009).
and mTOR pathways in biliary tract cancer. Br. J. Cancer 112, 1042–1051 445. Mendez, O. et al. Knock down of HIF-1alpha in glioma cells reduces migration
(2015). in vitro and invasion in vivo and impairs their ability to form tumor spheres. Mol.
415. Barker, A. et al. Validation of a non-targeted LC-MS approach for identifying Cancer 9, 133 (2010).
ancient proteins: method development on bone to improve artifact residue 446. Seidel, S. et al. A hypoxic niche regulates glioblastoma stem cells through
analysis. Ethnobiol. Lett. 6, 162–174 (2015). hypoxia inducible factor 2 alpha. Brain 133, 983–995 (2010).
416. Sturtzel, C. Endothelial cells. Adv. Exp. Med. Biol. 1003, 71–91 (2017). 447. Ye, X. Q. et al. Mitochondrial and energy metabolism-related properties as novel
417. Calabrese, C. et al. A perivascular niche for brain tumor stem cells. Cancer Cell indicators of lung cancer stem cells. Int. J. Cancer 129, 820–831 (2011).
11, 69–82 (2007). 448. Civenni, G. et al. RNAi-mediated silencing of Myc transcription inhibits stem-like
418. Beck, B. et al. A vascular niche and a VEGF-Nrp1 loop regulate the initiation and cell maintenance and tumorigenicity in prostate cancer. Cancer Res. 73,
stemness of skin tumours. Nature 478, 399–403 (2011). 6816–6827 (2013).
419. Lu, J. et al. Endothelial cells promote the colorectal cancer stem cell phenotype 449. Takebe, N., Warren, R. Q. & Ivy, S. P. Breast cancer growth and metastasis:
through a soluble form of Jagged-1. Cancer Cell 23, 171–185 (2013). interplay between cancer stem cells, embryonic signaling pathways and
420. Rong, W. et al. Glioblastoma stem-like cells give rise to tumour endothelium. epithelial-to-mesenchymal transition. Breast Cancer Res. 13, 211 (2011).
Nature 468, 829–833 (2010). 450. Bhat, K. M. Notch signaling acts before cell division to promote asymmetric
421. Scully, S. et al. Transdifferentiation of glioblastoma stem-like cells into mural cleavage and cell fate of neural precursor cells. Sci. Signal. 7, ra101 (2014).
cells drives vasculogenic mimicry in glioblastomas. J. Neurosci. 32, 12950–12960 451. Lavin, Y. et al. Tissue-resident macrophage enhancer landscapes are shaped by
(2012). the local microenvironment. Cell 159, 1312–1326 (2014).
422. Liu, T. J. et al. CD133+ cells with cancer stem cell characteristics associates with 452. Davies, L. C., Jenkins, S. J., Allen, J. E. & Taylor, P. R. Tissue-resident macrophages.
vasculogenic mimicry in triple-negative breast cancer. Oncogene 32, 544–553 Nat. Immunol. 14, 986–995 (2013).
(2013). 453. Sica, A. & Mantovani, A. Macrophage plasticity and polarization: in vivo veritas. J.
423. Yan, G. N. et al. Endothelial cells promote stem‐like phenotype of glioma cells Clin. Invest 122, 787–795 (2012).
through activating the Hedgehog pathway. J. Pathol. 234, 11–22 (2015). 454. Hao, N. B. et al. Macrophages in tumor microenvironments and the progression
424. Bao, S. et al. Stem cell-like glioma cells promote tumor angiogenesis through of tumors. Clin. Dev. Immunol. 2012, 948098 (2012).
vascular endothelial growth factor. Cancer Res. 66, 7843–7848 (2006). 455. Medrek, C., Ponten, F., Jirstrom, K. & Leandersson, K. The presence of tumor
425. Kaidi, A., Williams, A. C. & Paraskeva, C. Interaction between beta-catenin and associated macrophages in tumor stroma as a prognostic marker for breast
HIF-1 promotes cellular adaptation to hypoxia. Nat. Cell Biol. 9, 210–217 cancer patients. BMC Cancer 12, 306 (2012).
(2007). 456. Yamaguchi, T. et al. Tumor-associated macrophages of the M2 phenotype
426. Grange, C. et al. Microvesicles released from human renal cancer stem cells contribute to progression in gastric cancer with peritoneal dissemination.
stimulate angiogenesis and formation of lung premetastatic niche. Cancer Res. Gastric Cancer 19, 1052–1065 (2016).
71, 5346 (2011). 457. Zhou, P. et al. The epithelial to mesenchymal transition (EMT) and cancer stem
427. Bussolati, B., Grange, C., Sapino, A. & Camussi, G. Endothelial cell differentiation cells: implication for treatment resistance in pancreatic cancer. Mol. Cancer 16,
of human breast tumour stem/progenitor cells. J. Cell Mol. Med. 13, 309–319 52 (2017).
(2009). 458. Yang, Z., Xie, H., He, D. & Li, L. Infiltrating macrophages increase RCC epithelial
428. Alvero, A. B. et al. Stem-like ovarian cancer cells can serve as tumor vascular mesenchymal transition (EMT) and stem cell-like populations via AKT and mTOR
progenitors. Stem Cells 27, 2405–2413 (2009). signaling. Oncotarget 7, 44478–44491 (2016).
429. Pezzolo, A. et al. Oct-4+/Tenascin C+ neuroblastoma cells serve as progenitors 459. Huang, W. C., Chan, M. L., Chen, M. J., Tsai, T. H. & Chen, Y. J. Modulation of
of tumor-derived endothelial cells. Cell Res. 21, 1470–1486 (2011). macrophage polarization and lung cancer cell stemness by MUC1 and devel-
430. Zhao, Y. et al. Endothelial cell transdifferentiation of human glioma stem pro- opment of a related small-molecule inhibitor pterostilbene. Oncotarget 7,
genitor cells in vitro. Brain Res. Bull. 82, 308–312 (2010). 39363–39375 (2016).
431. Ping, Y. F., Zhang, X. & Bian, X. W. Cancer stem cells and their vascular niche: Do 460. Masahisa, J. et al. Tumor-associated macrophages regulate tumorigenicity and
they benefit from each other? Cancer Lett. 380, 561–567 (2016). anticancer drug responses of cancer stem/initiating cells. Proc. Natl Acad. Sci.
432. Krishnamurthy, S. et al. Endothelial cell-initiated signaling promotes the survival USA 108, 12425–12430 (2011).
and self-renewal of cancer stem cells. Cancer Res. 70, 9969–9978 (2010). 461. Wu, A. et al. Glioma cancer stem cells induce immunosuppressive macrophages/
433. Jennifer, P. et al. Microparticles mediated cross-talk between tumoral and microglia. Neuro Oncol. 12, 1113–1125 (2010).
endothelial cells promote the constitution of a pro-metastatic vascular niche 462. Yi, L. et al. Glioma-initiating cells: a predominant role in microglia/macrophages
through Arf6 up regulation. Cancer Microenviron. 7, 41–59 (2014). tropism to glioma. J. Neuroimmunol. 232, 75–82 (2011).
434. Lyden, D. et al. Impaired recruitment of bone-marrow-derived endothelial and 463. Yamashina, T. et al. Cancer stem-like cells derived from chemoresistant tumors
hematopoietic precursor cells blocks tumor angiogenesis and growth. Nat. Med. have a unique capacity to prime tumorigenic myeloid cells. Cancer Res. 74,
7, 1194–1201 (2001). 2698–2709 (2014).
435. Yao, X. et al. Vascular endothelial growth factor receptor 2 (VEGFR-2) plays a key 464. Leslie, C. S. CSF-1R inhibition alters macrophage polarization and blocks glioma
role in vasculogenic mimicry formation, neovascularization and tumor initiation progression. Nat. Med. 19, 1264 (2013).
by glioma stem-like cells. PLoS ONE 8, e57188- (2013). 465. Shi, Y. et al. Tumour-associated macrophages secrete pleiotrophin to promote
436. Galan-Moya, E. M. et al. Endothelial secreted factors suppress mitogen PTPRZ1 signalling in glioblastoma stem cells for tumour growth. Nat. Commun.
deprivation-induced autophagy and apoptosis in glioblastoma stem-like cells. 8, 15080 (2017).
PLoS ONE 9, e93505 (2014). 466. Xie, K. Interleukin-8 and human cancer biology. Cytokine Growth Factor Rev. 12,
437. Gilbertson, R. J. & Rich, J. N. Making a tumour's bed: glioblastoma stem cells and 375–391 (2001).
the vascular niche. Nat. Rev. Cancer 7, 733–736 (2007). 467. Yang, J. et al. Tumor-associated macrophages regulate murine breast cancer
438. Bao, S. et al. Glioma stem cells promote radioresistance by preferential activa- stem cells through a novel paracrine EGFR/Stat3/Sox-2 signaling pathway. Stem
tion of the DNA damage response. Nature 444, 756–760 (2006). Cells 31, 248–258 (2013).
439. Li, Z. et al. Hypoxia-inducible factors regulate tumorigenic capacity of glioma 468. Fu, X. T. et al. Macrophage-secreted IL-8 induces epithelial–mesenchymal
stem cells. Cancer Cell 15, 501–513 (2009). transition in hepatocellular carcinoma cells by activating the JAK2/STAT3/Snail
440. Kim, R. J. et al. High aldehyde dehydrogenase activity enhances stem cell fea- pathway. Int. J. Oncol. 46, 587–596 (2015).
tures in breast cancer cells by activating hypoxia-inducible factor-2alpha. Cancer 469. Wang, H. et al. HepG2 cells acquire stem cell-like characteristics after immune
Lett. 333, 18–31 (2013). cell stimulation. Cell Oncol. (Dordr.) 39, 35–45 (2016).
441. Keith, B. & Simon, M. C. Hypoxia-inducible factors, stem cells cancer. Cell 129, 470. Kalluri, R. & Zeisberg, M. Fibroblasts in cancer. Nat. Rev. Cancer 6, 392–401
465–472 (2007). (2006).
442. Hur, H. et al. Expression of pyruvate dehydrogenase kinase-1 in gastric cancer as 471. Potenta, S., Zeisberg, E. & Kalluri, R. The role of endothelial-to-mesenchymal
a potential therapeutic target. Int. J. Oncol. 42, 44–54 (2013). transition in cancer progression. Br. J. Cancer 99, 1375–1379 (2008).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
30
472. Crisan, M. et al. A perivascular origin for mesenchymal stem cells in multiple 501. Tu, B., L, D., Fan, Q. M., Tang, Z. & Tang, T. T. STAT3 activation by IL-6 from
human organs. Cell Stem Cell 3, 301–313 (2008). mesenchymal stem cells promotes the proliferation and metastasis of osteo-
473. Kalluri, R. & Weinberg, R. A. The basics of epithelial–mesenchymal transition. J. sarcoma. Cancer Lett. 325, 80–88 (2012).
Clin. Invest 119, 1420–1428 (2009). 502. Spaeth, E. L. et al. Mesenchymal stem cell transition to tumor-associated
474. Buchsbaum, R. J. & Young, O. S. Breast cancer-associated fibroblasts: where we fibroblasts contributes to fibrovascular network expansion and tumor progres-
are and where we need to go. Cancers 8, 19 (2016). sion. PLoS ONE 8, e4992 (2013).
475. Jotzu, C. et al. Adipose tissue-derived stem cells differentiate into carcinoma- 503. Reagan, M. R. & Kaplan, D. L. Concise review: mesenchymal stem cell tumor-
associated fibroblast-like cells under the influence of tumor-derived factors. Cell. homing: detection methods in disease model systems. Stem Cells 29, 920–927
Oncol. 34, 55–67 (2011). (2011).
476. Wikstrom, P., Marusic, J., Stattin, P. & Bergh, A. Low stroma androgen receptor 504. Yang, Y., Otte, A. & Hass, R. Human mesenchymal stroma/stem cells exchange
level in normal and tumor prostate tissue is related to poor outcome in prostate membrane proteins and alter functionality during interaction with different
cancer patients. Prostate 69, 799–809 (2009). tumor cell lines. Stem Cells Dev. 24, 1205–1222 (2014).
477. Goicoechea, S. M. et al. Palladin promotes invasion of pancreatic cancer cells by 505. Xu, M. H. et al. EMT and acquisition of stem cell-like properties are involved in
enhancing invadopodia formation in cancer-associated fibroblasts. Oncogene spontaneous formation of tumorigenic hybrids between lung cancer and bone
33, 1265–1273 (2014). marrow-derived mesenchymal stem cells. PLoS ONE 9, e87893 (2014).
478. Xia, L. et al. A CCL2/ROS autoregulation loop is critical for cancer-associated 506. Yang, X. et al. Increased invasiveness of osteosarcoma mesenchymal stem cells
fibroblasts-enhanced tumor growth of oral squamous cell carcinoma. Carcino- induced by bone-morphogenetic protein-2. Vitr. Cell Dev. Biol. Anim. 49,
genesis 35, 1362–1370 (2014). 270–278 (2013).
479. Nair, N. et al. A cancer stem cell model as the point of origin of cancer- 507. Ma, Y. C. et al. The tyrosine kinase c-Src directly mediates growth factor-induced
associated fibroblasts in tumor microenvironment. Sci. Rep. 7, 6838 (2017). Notch-1 and Furin interaction and Notch-1 activation in pancreatic cancer cells.
480. Scheel, C. et al. Paracrine and autocrine signals induce and maintain PLoS ONE 7, e33414 (2012).
mesenchymal and stem cell states in the breast. Cell 145, 926–940 (2011). 508. Carnero, A. et al. The cancer stem-cell signaling network and resistance to
481. Giannoni, E. et al. Reciprocal activation of prostate cancer cells and cancer- therapy. Cancer Treat. Rev. 49, 25–36 (2016).
associated fibroblasts stimulates epithelial–mesenchymal transition and cancer 509. Peng, D. et al. Myeloid-derived suppressor cells endow stem-like qualities to
stemness. Cancer Res. 70, 6945–6956 (2010). breast cancer cells through IL6/STAT3 and NO/NOTCH cross-talk signaling.
482. Liao, C. P., Adisetiyo, H., Liang, M. & Roy-Burman, P. Cancer-associated fibroblasts Cancer Res. 76, 3156 (2016).
enhance the gland-forming capability of prostate cancer stem cells. Cancer Res. 510. Hynes, R. O. & Naba, A. Overview of the matrisome—an inventory of extra-
70, 7294–7303 (2010). cellular matrix constituents and functions. Cold Spring Harb. Perspect. Biol. 4,
483. Wolfson, B., Eades, G. & Zhou, Q. Adipocyte activation of cancer stem cell sig- a004903 (2012).
naling in breast cancer. World J. Biol. Chem. 6, 39–47 (2015). 511. Mouw, J. K., Ou, G. & Weaver, V. M. Extracellular matrix assembly: a multiscale
484. Chaffer, C. L. et al. Poised chromatin at the ZEB1 promoter enables cell plasticity deconstruction. Nat. Rev. Mol. Cell Biol. 15, 771–785 (2014).
and enhances tumorigenicity. Cell 154, 61–74 (2013). 512. Pengfei, L., Ken, T., Weaver, V. M. & Zena, W. Extracellular matrix degradation
485. Chen, W. J. et al. Cancer-associated fibroblasts regulate the plasticity of lung and remodeling in development and disease. Cold Spring Harbor Perspectives in.
cancer stemness via paracrine signalling. Nat. Commun. 5, 3472 (2014). Cold Spring Harbor Perspect. Biol. 3, 1750–1754 (2011).
486. Tsuyada, A. et al. CCL2 mediates cross-talk between cancer cells and stromal 513. Pengfei, L., Weaver, V. M. & Zena, W. The extracellular matrix: a dynamic niche in
fibroblasts that regulates breast cancer stem cells. Cancer Res. 72, 2768 cancer progression. J. Cell Biol. 196, 395 (2012).
(2012). 514. Li, L., John, C. & Margolin, D. A. Cancer stem cell and stromal microenvironment.
487. Giordano, C. et al. Leptin as a mediator of tumor-stromal interactions promotes Ochsner J. 13, 109–118 (2013).
breast cancer stem cell activity. Oncotarget 7, 1262–1275 (2016). 515. Schrader, J. et al. Matrix stiffness modulates proliferation, chemotherapeutic
488. Mccuaig, R., Wu, F., Dunn, J., Rao, S. & Dahlstrom, J. E. The biological and clinical response, and dormancy in hepatocellular carcinoma cells. Hepatology 53,
significance of stromal-epithelial interactions in breast cancer. Pathology 49, 1192–1205 (2011).
133–140 (2016). 516. Casazza, A. et al. Tumor stroma: a complexity dictated by the hypoxic tumor
489. Lau, E. Y. et al. Cancer-associated fibroblasts regulate tumor-initiating cell microenvironment. Oncogene 33, 1743–1754 (2014).
plasticity in hepatocellular carcinoma through c-Met/FRA1/HEY1 signaling. Cell 517. Murai, T. Lipid raft-mediated regulation of hyaluronan–CD44 interactions in
Rep. 15, 1175–1189 (2016). inflammation and cancer. Front. Immunol. 6, 420 (2015).
490. Xiong, S. et al. Cancer-associated fibroblasts promote stem cell-like properties of 518. Kai, K. et al. A role for matrix metalloproteinases in regulating mammary stem
hepatocellular carcinoma cells through IL-6/STAT3/Notch signaling. Am. J. cell function via the Wnt signaling pathway. Cell Stem Cell 13, 300–313 (2013).
Cancer Res. 8, 302 (2018). 519. Thordur, O. et al. Breast cancer cells produce tenascin C as a metastatic niche
491. Yibing, H. et al. Fibroblast-derived exosomes contribute to chemoresistance component to colonize the lungs. Nat. Med. 17, 867–874 (2011).
through priming cancer stem cells in colorectal cancer. PLoS ONE 10, e0125625 520. Yuan, A. et al. Transfer of microRNAs by embryonic stem cell microvesicles. PLoS
(2015). ONE 4, e4722 (2009).
492. Bagley, R. G. et al. Human mesenchymal stem cells from bone marrow express 521. Vlassov, A. V., Susan, M., Robert, S. & Rick, C. Exosomes: current knowledge of
tumor endothelial and stromal markers. Int. J. Oncol. 34, 619–627 (2009). their composition, biological functions, and diagnostic and therapeutic poten-
493. Cavarretta, I. T. et al. Adipose tissue-derived mesenchymal stem cells expressing tials. Biochim. Biophys. Acta 1820, 940–948 (2012).
prodrug-converting enzyme inhibit human prostate tumor growth. Mol. Ther. 522. Kawikova, I. & Askenase, P. W. Diagnostic and therapeutic potentials of exo-
18, 223–231 (2010). somes in CNS diseases. Brain Res. 1617, 63–71 (2015).
494. Kidd, S. et al. Direct evidence of mesenchymal stem cell tropism for tumor and 523. Pete, H. et al. A family of thermostable fungal cellulases created by structure-
wounding microenvironments using in vivo bioluminescent imaging. Stem Cells guided recombination. Proc. Natl Acad. Sci. USA 106, 5610–5615 (2009).
27, 2614–2623 (2009). 524. Alonso, M. A. & Millán, J. The role of lipid rafts in signalling and membrane
495. Martin, F. T. et al. Potential role of mesenchymal stem cells (MSCs) in the breast trafficking in T lymphocytes. J. Cell Sci. 114, 3957 (2001).
tumour microenvironment: stimulation of epithelial to mesenchymal transition 525. Chow, A. et al. Macrophage immunomodulation by breast cancer-derived
(EMT). Breast Cancer Res. Treat. 124, 317–326 (2010). exosomes requires Toll-like receptor 2-mediated activation of NF-κB. Sci. Rep. 4,
496. Xue, J. et al. Tumorigenic hybrids between mesenchymal stem cells and gastric 5750 (2014).
cancer cells enhanced cancer proliferation, migration and stemness. BMC Cancer 526. Lopatina, T., Gai, C., Deregibus, M. C., Kholia, S. & Camussi, G. Cross talk between
15, 793 (2015). cancer and mesenchymal stem cells through extracellular vesicles carrying
497. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell nucleic acids. Front. Oncol. 6, 125 (2016).
144, 646–674 (2011). 527. Abels, E. R. & Breakefield, X. O. Introduction to extracellular vesicles: biogenesis,
498. Brunel, M., Herr, F. & Durrbach, A. Immunosuppressive properties of mesench- RNA cargo selection, content, release, and uptake. Cell. Mol. Neurobiol. 36,
ymal stem cells. Curr. Transplant. Rep. 3, 348–357 (2016). 301–312 (2016).
499. Abdel, A. M. T. et al. Efficacy of mesenchymal stem cells in suppression of 528. Gajos-Michniewicz, A., Duechler, M. & Czyz, M. MiRNA in melanoma-derived
hepatocarcinorigenesis in rats: possible role of wnt signaling. J. Exp. Clin. Cancer exosomes. Cancer Lett. 347, 29–37 (2014).
Res. 30, 49–49 (2011). 529. Webber, J. P. et al. Differentiation of tumour-promoting stromal myofibroblasts
500. Maffey, A. et al. Mesenchymal stem cells from tumor microenvironment favour by cancer exosomes. Oncogene 34, 290 (2015).
breast cancer stem cell proliferation, cancerogenic and metastatic potential, via 530. Mrizak, D. et al. 19 Effect of nasopharyngeal carcinoma-derived exosomes on
ionotropic purinergic signalling. Sci. Rep. 7, 13162 (2017). human regulatory T cells. J. Natl Cancer Inst. 51, e33–e33 (2015).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
31
531. Rana, S., Malinowska, K. & Zöller, M. Exosomal tumor microRNA modulates 558. Pemmaraju, N. et al. Results from phase 2 trial ongoing expansion stage of SL-
premetastatic organ cells 1 2. Neoplasia 15, 281 (2013). IN214–IN295, IN231. 401 in patients with blastic plasmacytoid dendritic cell neoplasm (BPDCN).
532. Melo, S. et al. Cancer exosomes perform cell-independent microRNA biogenesis Blood 128, 342–342 (2016).
and promote tumorigenesis. Cancer Cell 26, 707–721 (2014). 559. Osta, W. A. et al. EpCAM is overexpressed in breast cancer and is a potential
533. Andrew, R., Elaine, C., Eva, B. & Simon, P. Regulation of exosome release from target for breast cancer gene therapy. Cancer Res. 64, 5818–5824 (2004).
mammary epithelial and breast cancer cells—a new regulatory pathway. Eur. J. 560. Oberneder, R. et al. A phase I study with adecatumumab, a human antibody
Cancer 50, 1025–1034 (2014). directed against epithelial cell adhesion molecule, in hormone refractory
534. Nakano, I., Garnier, D., Minata, M. & Rak, J. Extracellular vesicles in the biology of prostate cancer patients. Eur. J. Cancer 42, 2530–2538 (2006).
brain tumour stem cells—implications for inter-cellular communication, therapy 561. Sebastian, M. et al. Treatment of non-small cell lung cancer patients with the
and biomarker development. Semin. Cell Dev. Biol. 40, 17–26 (2015). trifunctional monoclonal antibody catumaxomab (anti-EpCAM × anti-CD3): a
535. Rinkenbaugh, A. L. & Baldwin, A. S. The NF-κB pathway and cancer stem cells. phase I study. Cancer Immunol. Immunother. 56, 1637–1644 (2007).
Cells 5, 16 (2016). 562. Jaeger, M. et al. Immunotherapy with the trifunctional antibody removab leads
536. Dandawate, P. R., Subramaniam, D., Jensen, R. A. & Anant, S. Targeting cancer to significant elimination of tumor cells from malignant ascites in ovarian
stem cells and signaling pathways by phytochemicals: novel approach for cancer: results of a phase I/II study. J. Clin. Oncol. 22, 2504–2504 (2004).
breast cancer therapy. Semin. Cancer Biol. 40–41, 192–208 (2016). 563. Niedzwiecki, D. et al. Documenting the natural history of patients with resected
537. Huang, R. & Rofstad, E. K. Cancer stem cells (CSCs), cervical CSCs and targeted stage II adenocarcinoma of the colon after random assignment to adjuvant
therapies. Oncotarget 8, 35351–35367 (2017). treatment with edrecolomab or observation: results from CALGB 9581. J. Clin.
538. Bronisz, A. et al. Extracellular vesicles modulate the glioblastoma micro- Oncol. 29, 3146–3152 (2011).
environment via a tumor suppression signaling network directed by miR-1. 564. Schmidt, M. et al. An open-label, randomized phase II study of adecatumumab,
Cancer Res. 74, 738–750 (2014). a fully human anti-EpCAM antibody, as monotherapy in patients with metastatic
539. Ye, J., Wu, D., Wu, P., Chen, Z. & Huang, J. The cancer stem cell niche: cross talk breast cancer. Ann. Oncol. 21, 275–282 (2010).
between cancer stem cells and their microenvironment. Tumor Biol. 35, 565. Quail, D. F., Taylor, M. J. & Postovit, L. M. Microenvironmental regulation of
3945–3951 (2014). cancer stem cell phenotypes. Curr. Stem Cell Res Ther. 7, 197–216 (2012).
540. Shimoda, M. et al. Loss of the Timp gene family is sufficient for the acquisition of 566. Meurette, O. & Mehlen, P. Notch signaling in the tumor microenvironment.
the CAF-like cell state. Nat. Cell Biol. 16, 889–901 (2014). Cancer Cell 34, 536–548 (2018).
541. Donnarumma, E. et al. Cancer-associated fibroblasts release exosomal micro- 567. Nowell, C. S. & Radtke, F. Notch as a tumour suppressor. Nat. Rev. Cancer 17,
RNAs that dictate an aggressive phenotype in breast cancer. Oncotarget 8, 145–159 (2017).
19592–19608 (2017). 568. Rajakulendran, N. et al. Wnt and Notch signaling govern self-renewal and dif-
542. Figueroa, J. et al. Exosomes from glioma-associated mesenchymal stem cells ferentiation in a subset of human glioblastoma stem cells. Genes Dev. 33,
increase the tumorigenicity of glioma stem-like cells via transfer of miR-1587. 498–510 (2019).
Cancer Res. 77, 5808 (2017). 569. van Groningen, T. et al. A NOTCH feed-forward loop drives reprogramming from
543. Xu, J., Liao, K. & Zhou, W. Exosomes regulate the transformation of cancer cells adrenergic to mesenchymal state in neuroblastoma. Nat. Commun. 10, 1530
in cancer stem cell homeostasis. Stem Cells Int. 2018, 4837370 (2018). (2019).
544. Munagala, R., Aqil, F., Jeyabalan, J. & Gupta, R. C. Bovine milk-derived exosomes 570. Zanotti, S. & Canalis, E. Notch signaling and the skeleton. Endocr. Rev. 37,
for drug delivery. Cancer Lett. 371, 48–61 (2016). 223–253 (2016).
545. Srivastava, A. et al. Exploitation of exosomes as nanocarriers for gene-, chemo-, 571. Gazdar, A. F., Bunn, P. A. & Minna, J. D. Small-cell lung cancer: what we know,
and immune-therapy of cancer. J. Biomed. Nanotechnol. 12, 1159 (2016). what we need to know and the path forward. Nat. Rev. Cancer 17, 725–737
546. Pitt, J. M. et al. Dendritic cell-derived exosomes for cancer therapy. J. Clin. Invest. (2017).
126, 1224 (2016). 572. Dai, W., Peterson, A., Kenney, T., Burrous, H. & Montell, D. J. Quantitative
547. Tian, X., Zhu, M. & Nie, G. How can nanotechnology help membrane vesicle- microscopy of the Drosophila ovary shows multiple niche signals specify pro-
based cancer immunotherapy development? Hum. Vaccin. Immunother. 9, genitor cell fate. Nat. Commun. 8, 1244 (2017).
222–225 (2013). 573. Mamidi, A. et al. Mechanosignalling via integrins directs fate decisions of pan-
548. Wang, J. et al. Extracellular vesicle cross-talk in the bone marrow micro- creatic progenitors. Nature 564, 114–118 (2018).
environment: implications in multiple myeloma. Oncotarget 7, 38927–38945 574. Hayakawa, Y. et al. BHLHA15-positive secretory precursor cells can give rise to
(2016). tumors in intestine and colon in mice. Gastroenterology 156, 1066–1081 (2019).
549. Ghielmini, M. et al. The effect of Rituximab on patients with follicular and e1016.
mantle-cell lymphoma. Swiss Group for Clinical Cancer Research (SAKK). Ann. 575. Fouladi, M. et al. Phase I trial of MK-0752 in children with refractory CNS
Oncol. 11(Suppl. 1), 123–126 (2000). malignancies: a pediatric brain tumor consortium study. J. Clin. Oncol. 29,
550. Turner, J. H., Martindale, A. A., Boucek, J., Claringbold, P. G. & Leahy, M. F. 131I- 3529–3534 (2011).
anti CD20 radioimmunotherapy of relapsed or refractory non-Hodgkins lym- 576. Krop, I. et al. Phase I pharmacologic and pharmacodynamic study of the gamma
phoma: a phase II clinical trial of a nonmyeloablative dose regimen of chimeric secretase (Notch) inhibitor MK-0752 in adult patients with advanced solid
rituximab radiolabeled in a hospital. Cancer Biother. Radiopharm. 18, 513–524 tumors. J. Clin. Oncol. 30, 2307–2313 (2012).
(2003). 577. Hoffman, L. M. et al. Phase I trial of weekly MK-0752 in children with refractory
551. Nabhan, C. et al. A pilot trial of rituximab and alemtuzumab combination central nervous system malignancies: a pediatric brain tumor consortium study.
therapy in patients with relapsed and/or refractory chronic lymphocytic leuke- Childs Nerv. Syst. 31, 1283–1289 (2015).
mia (CLL). Leuk. Lymphoma 45, 2269–2273 (2004). 578. Chen, X. et al. Sequential combination therapy of ovarian cancer with cisplatin
552. Börjesson, P. K. et al. Phase I therapy study with (186)Re-labeled humanized and γ-secretase inhibitor MK-0752. Gynecol. Oncol. 140, 537–544 (2016).
monoclonal antibody BIWA 4 (bivatuzumab) in patients with head and neck 579. Schott, A. F. et al. Preclinical and clinical studies of gamma secretase inhibitors
squamous cell carcinoma. Clin. Cancer Res. 9, 3961s–3972s (2003). with docetaxel on human breast tumors. Clin. Cancer Res. 19, 1512–1524 (2013).
553. Postema, E. J. et al. Dosimetric analysis of radioimmunotherapy with 186Re- 580. Cook, N. et al. A phase I trial of the γ-secretase inhibitor MK-0752 in combination
labeled bivatuzumab in patients with head and neck cancer. J. Nucl. Med. 44, with gemcitabine in patients with pancreatic ductal adenocarcinoma. Br. J.
1690–1699 (2003). Cancer 118, 793–801 (2018).
554. Colnot, D. R. et al. Safety, biodistribution, pharmacokinetics, and immunogeni- 581. Brana, I. et al. A parallel-arm phase I trial of the humanised anti-IGF-1R antibody
city of 99mTc-labeled humanized monoclonal antibody BIWA 4 (bivatuzumab) dalotuzumab in combination with the AKT inhibitor MK-2206, the mTOR inhi-
in patients with squamous cell carcinoma of the head and neck. Cancer bitor ridaforolimus, or the NOTCH inhibitor MK-0752, in patients with advanced
Immunol. Immunother. 52, 576–582 (2003). solid tumours. Br. J. Cancer 111, 1932–1944 (2014).
555. Riechelmann, H. et al. Phase I trial with the CD44v6-targeting immunoconjugate 582. Zhang, S., Chung, W. C., Miele, L. & Xu, K. Targeting Met and Notch in the Lfng-
bivatuzumab mertansine in head and neck squamous cell carcinoma. Oral. deficient, Met-amplified triple-negative breast cancer. Cancer Biol. Ther. 15,
Oncol. 44, 823–829 (2008). 633–642 (2014).
556. He, S. Z. et al. A phase 1 study of the safety, pharmacokinetics and anti-leukemic 583. Luistro, L. et al. Preclinical profile of a potent gamma-secretase inhibitor tar-
activity of the anti-CD123 monoclonal antibody CSL360 in relapsed, refractory geting notch signaling with in vivo efficacy and pharmacodynamic properties.
or high-risk acute myeloid leukemia. Leuk. Lymphoma 56, 1406–1415 (2015). Cancer Res. 69, 7672–7680 (2009).
557. Frankel, A. E. et al. Activity of SL-401, a targeted therapy directed to interleukin-3 584. Tolcher, A. W. et al. Phase I study of RO4929097, a gamma secretase inhibitor of
receptor, in blastic plasmacytoid dendritic cell neoplasm patients. Blood 124, Notch signaling, in patients with refractory metastatic or locally advanced solid
385–392 (2014). tumors. J. Clin. Oncol. 30, 2348–2353 (2012).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
32
585. Strosberg, J. R. et al. A phase II study of RO4929097 in metastatic colorectal 612. Kieran, M. W. et al. Phase I study of oral sonidegib (LDE225) in pediatric brain
cancer. Eur. J. Cancer 48, 997–1003 (2012). and solid tumors and a phase II study in children and adults with relapsed
586. De Jesus-Acosta, A. et al. A phase II study of the gamma secretase inhibitor medulloblastoma. Neuro Oncol. 19, 1542–1552 (2017).
RO4929097 in patients with previously treated metastatic pancreatic adeno- 613. Magnani, L. et al. Genome-wide reprogramming of the chromatin landscape
carcinoma. Invest. N. Drugs 32, 739–745 (2014). underlies endocrine therapy resistance in breast cancer. Proc. Natl Acad. Sci. USA
587. Diaz-Padilla, I. et al. A phase II study of single-agent RO4929097, a gamma- 110, E1490–E1499 (2013).
secretase inhibitor of Notch signaling, in patients with recurrent platinum- 614. Cortes, J. E. et al. Glasdegib in combination with cytarabine and daunorubicin in
resistant epithelial ovarian cancer: a study of the Princess Margaret, Chicago and patients with AML or high-risk MDS: phase 2 study results. Am. J. Hematol. 93,
California phase II consortia. Gynecol. Oncol. 137, 216–222 (2015). 1301–1310 (2018).
588. Richter, S. et al. A phase I study of the oral gamma secretase inhibitor 615. Cortes, J. E. et al. Randomized comparison of low dose cytarabine with or
R04929097 in combination with gemcitabine in patients with advanced solid without glasdegib in patients with newly diagnosed acute myeloid leukemia or
tumors (PHL-078/CTEP 8575). Invest N. Drugs 32, 243–249 (2014). high-risk myelodysplastic syndrome. Leukemia 33, 379–389 (2019).
589. Sahebjam, S. et al. A phase I study of the combination of ro4929097 and 616. List, A. et al. Opportunities for Trisenox (arsenic trioxide) in the treatment of
cediranib in patients with advanced solid tumours (PJC-004/NCI 8503). Br. J. myelodysplastic syndromes. Leukemia 17, 1499–1507 (2003).
Cancer 109, 943–949 (2013). 617. Lauth, M., Bergström, A., Shimokawa, T. & Toftgård, R. Inhibition of GLI-mediated
590. LoConte, N. K. et al. A multicenter phase 1 study of γ -secretase inhibitor transcription and tumor cell growth by small-molecule antagonists. Proc. Natl
RO4929097 in combination with capecitabine in refractory solid tumors. Invest. Acad. Sci. USA 104, 8455–8460 (2007).
N. Drugs 33, 169–176 (2015). 618. Pohl, S. G. et al. Wnt signaling in triple-negative breast cancer. Oncogenesis 6,
591. Kummar, S. et al. Clinical activity of the γ-secretase inhibitor PF-03084014 in e310 (2017).
adults with desmoid tumors (aggressive fibromatosis). J. Clin. Oncol. 35, 619. Kraggerud, S. M. et al. Molecular characteristics of malignant ovarian germ cell
1561–1569 (2017). tumors and comparison with testicular counterparts: implications for patho-
592. Messersmith, W. A. et al. A phase I, dose-finding study in patients with advanced genesis. Endocr. Rev. 34, 339–376 (2013).
solid malignancies of the oral γ-secretase inhibitor PF-03084014. Clin. Cancer 620. Fu, L. et al. Wnt2 secreted by tumour fibroblasts promotes tumour progression
Res. 21, 60–67 (2015). in oesophageal cancer by activation of the Wnt/β-catenin signalling pathway.
593. Papayannidis, C. et al. A phase 1 study of the novel gamma-secretase inhibitor Gut 60, 1635–1643 (2011).
PF-03084014 in patients with T-cell acute lymphoblastic leukemia and T-cell 621. Hua, F. et al. TRIB3 interacts with β-catenin and TCF4 to increase stem cell
lymphoblastic lymphoma. Blood Cancer J. 5, e350 (2015). features of colorectal cancer stem cells and tumorigenesis. Gastroenterology
594. Yang, J. et al. Role of Jagged1/STAT3 signalling in platinum-resistant ovarian 156, 708–721 (2019). e715.
cancer. J. Cell. Mol. Med. 23, 4005–4018 (2019). 622. Pal, S. K., Swami, U. & Agarwal, N. Characterizing the Wnt pathway in advanced
595. Smith, D. C. et al. A phase I dose escalation and expansion study of the antic- prostate cancer: when, why, and how. Eur. Urol. (2019).
ancer stem cell agent demcizumab (anti-DLL4) in patients with previously 623. Lin, W. et al. Mesenchymal stem cells and cancer: clinical challenges and
treated solid tumors. Clin. Cancer Res. 20, 6295–6303 (2014). opportunities. Biomed. Res. Int. 2019, 2820853 (2019).
596. Mukherjee, S. et al. Hedgehog signaling and response to cyclopamine differ in 624. Le, P. N., McDermott, J. D. & Jimeno, A. Targeting the Wnt pathway in human
epithelial and stromal cells in benign breast and breast cancer. Cancer Biol. Ther. cancers: therapeutic targeting with a focus on OMP-54F28. Pharmacol. Ther.
5, 674–683 (2006). 146, 1–11 (2015).
597. Yuan, Z. et al. Frequent requirement of hedgehog signaling in non-small cell 625. Jimeno, A. et al. Phase I study of the Hedgehog pathway inhibitor IPI-926 in
lung carcinoma. Oncogene 26, 1046–1055 (2007). adult patients with solid tumors. Clin. Cancer Res. 19, 2766–2774 (2013).
598. Thayer, S. P. et al. Hedgehog is an early and late mediator of pancreatic cancer 626. Cortes, J. E. et al. Phase 1 study of CWP232291 in relapsed/refractory acute
tumorigenesis. Nature 425, 851–856 (2003). myeloid leukemia (AML) and myelodysplastic syndrome (MDS). J. Clin. Oncol. 33,
599. Sekulic, A. et al. Efficacy and safety of vismodegib in advanced basal-cell car- 7044–7044 (2015).
cinoma. N. Engl. J. Med. 366, 2171–2179 (2012). 627. Konopleva, M. et al. Efficacy and biological correlates of response in a phase II
600. Dummer, R. et al. The 12-month analysis from basal cell carcinoma outcomes study of venetoclax monotherapy in patients with acute myelogenous leuke-
with LDE225 treatment (BOLT): a phase II, randomized, double-blind study of mia. Cancer Discov. 6, 1106–1117 (2016).
sonidegib in patients with advanced basal cell carcinoma. J. Am. Acad. Dermatol 628. Saygin, C., Matei, D., Majeti, R., Reizes, O. & Lathia, J. D. Targeting cancer
75, 113–125 (2016). e115. stemness in the clinic: from hype to hope. Cell Stem Cell 24, 25–40 (2019).
601. Norsworthy, K. J. et al. FDA approval summary: glasdegib for newly diagnosed 629. Cashen, A. et al. A phase II study of plerixafor (AMD3100) plus G-CSF for auto-
acute myeloid leukemia. Clin. Cancer Res. 25, 6021–6025 (2019). logous hematopoietic progenitor cell mobilization in patients with Hodgkin
602. Scales, S. J. & de Sauvage, F. J. Mechanisms of Hedgehog pathway activation in lymphoma. Biol. Blood Marrow Transplant. 14, 1253–1261 (2008).
cancer and implications for therapy. Trends Pharmacol. Sci. 30, 303–312 (2009). 630. Uy, G. L. et al. A phase 1/2 study of chemosensitization with the CXCR4
603. Zito, P. M. & Scharf, R. in StatPearls (StatPearls Publishing StatPearls Publishing antagonist plerixafor in relapsed or refractory acute myeloid leukemia. Blood
LLC., Treasure Island, FL, 2019). 119, 3917–3924 (2012).
604. Robinson, G. W. et al. Vismodegib exerts targeted efficacy against recurrent 631. Cooper, T. M. et al. A phase 1 study of the CXCR4 antagonist plerixafor in
sonic hedgehog-subgroup medulloblastoma: results from phase II pediatric combination with high-dose cytarabine and etoposide in children with relapsed
brain tumor consortium studies PBTC-025B and PBTC-032. J. Clin. Oncol. 33, or refractory acute leukemias or myelodysplastic syndrome: a Pediatric Oncol-
2646–2654 (2015). ogy Experimental Therapeutics Investigators' Consortium study (POE 10-03).
605. Gajjar, A. et al. Phase I study of vismodegib in children with recurrent or Pediatr. Blood Cancer 64, (2017).
refractory medulloblastoma: a pediatric brain tumor consortium study. Clin. 632. Galsky, M. D. et al. A phase I trial of LY2510924, a CXCR4 peptide antagonist, in
Cancer Res. 19, 6305–6312 (2013). patients with advanced cancer. Clin. Cancer Res. 20, 3581–3588 (2014).
606. Berlin, J. et al. A randomized phase II trial of vismodegib versus placebo with 633. Hainsworth, J. D. et al. A randomized, open-label phase 2 study of the CXCR4
FOLFOX or FOLFIRI and bevacizumab in patients with previously untreated inhibitor LY2510924 in combination with sunitinib versus sunitinib alone in
metastatic colorectal cancer. Clin. Cancer Res. 19, 258–267 (2013). patients with metastatic renal cell carcinoma (RCC). Target Oncol. 11, 643–653
607. Catenacci, D. V. et al. Randomized phase Ib/II study of gemcitabine plus placebo (2016).
or vismodegib, a hedgehog pathway inhibitor, in patients with metastatic 634. Salgia, R. et al. A randomized phase II study of LY2510924 and carboplatin/
pancreatic cancer. J. Clin. Oncol. 33, 4284–4292 (2015). etoposide versus carboplatin/etoposide in extensive-disease small cell lung
608. Italiano, A. et al. GDC-0449 in patients with advanced chondrosarcomas: a cancer. Lung Cancer 105, 7–13 (2017).
French Sarcoma Group/US and French National Cancer Institute Single-Arm 635. Sakamuri, D. et al. Phase I dose-escalation study of anti-CTLA-4 antibody ipili-
Phase II Collaborative Study. Ann. Oncol. 24, 2922–2926 (2013). mumab and lenalidomide in patients with advanced cancers. Mol. Cancer Ther.
609. Houot, R. et al. Inhibition of Hedgehog signaling for the treatment of lymphoma 17, 671–676 (2018).
and CLL: a phase II study from the LYSA. Ann. Oncol. 27, 1349–1350 (2016). 636. Meindl-Beinker, N. M. et al. A multicenter open-label phase II trial to evaluate
610. Kaye, S. B. et al. A phase II, randomized, placebo-controlled study of vismodegib nivolumab and ipilimumab for 2nd line therapy in elderly patients with
as maintenance therapy in patients with ovarian cancer in second or third advanced esophageal squamous cell cancer (RAMONA). BMC Cancer 19, 231
complete remission. Clin. Cancer Res. 18, 6509–6518 (2012). (2019).
611. Pan, S. et al. Discovery of NVP-LDE225, a potent and selective smoothened 637. Cortese, I. et al. Pembrolizumab treatment for progressive multifocal leu-
antagonist. ACS Med. Chem. Lett. 1, 130–134 (2010). koencephalopathy. N. Engl. J. Med. 380, 1597–1605 (2019).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
33
638. Migden, M. R. et al. PD-1 blockade with cemiplimab in advanced cutaneous 670. Vassilopoulos, A., Chisholm, C., Lahusen, T., Zheng, H. & Deng, C. X. A critical role
squamous-cell carcinoma. N. Engl. J. Med. 379, 341–351 (2018). of CD29 and CD49f in mediating metastasis for cancer-initiating cells isolated
639. Motzer, R. J. et al. Avelumab plus Axitinib versus Sunitinib for advanced renal- from a Brca1-associated mouse model of breast cancer. Oncogene 33,
cell carcinoma. N. Engl. J. Med. 380, 1103–1115 (2019). 5477–5482 (2014).
640. Fujiwara, Y. et al. Tolerability and efficacy of durvalumab in Japanese patients 671. Deng, Z. et al. Adoptive T-cell therapy of prostate cancer targeting the cancer
with advanced solid tumors. Cancer Sci. 110, 1715–1723 (2019). stem cell antigen EpCAM. BMC immunology. 16, 1 (2015).
641. Sullivan, R. J. et al. Atezolizumab plus cobimetinib and vemurafenib in BRAF- 672. Kerr, B. A. et al. CD117(+) cells in the circulation are predictive of advanced
mutated melanoma patients. Nat. Med. 25, 929–935 (2019). prostate cancer. Oncotarget 6, 1889–1897 (2015).
642. Hodi, F. S. et al. Improved survival with ipilimumab in patients with metastatic 673. Patrawala, L. et al. Highly purified CD44+ prostate cancer cells from xenograft
melanoma. N. Engl. J. Med. 363, 711–723 (2010). human tumors are enriched in tumorigenic and metastatic progenitor cells.
643. Nandy, S. B. & Lakshmanaswamy, R. Cancer stem cells and metastasis. Prog. Mol. Oncogene 25, 1696–1708 (2006).
Biol. Transl. Sci. 151, 137–176 (2017). 674. Ugolkov, A. V., Eisengart, L. J., Luan, C. Y. & Yang, X. M. J. Expression analysis of
644. Rich, J. N. Cancer stem cells: understanding tumor hierarchy and heterogeneity. putative stem cell markers in human benign and malignant prostate. Prostate
Medicine 95, S2–S7 (2016). 71, 18–25 (2011).
645. Prieto-Vila, M., Takahashi, R. U., Usuba, W., Kohama, I. & Ochiya, T. Drug resis- 675. Darash-Yahana, M. et al. Role of high expression levels of CXCR4 in tumor
tance driven by cancer stem cells and their niche. Int. J. Mol. Sci. 18, 2574 (2017). growth, vascularization, and metastasis. FASEB J. 18, 1240-+ (2004).
646. Zhao, J. Cancer stem cells and chemoresistance: the smartest survives the raid. 676. Bae, K. M., Parker, N. N., Dai, Y., Vieweg, J. & Siemann, D. W. E-cadherin plasticity
Pharmacol. Ther. 160, 145–158 (2016). in prostate cancer stem cell invasion. Am. J. Cancer Res. 1, 71–84 (2011).
647. Chang, L. et al. Cancer stem cells and signaling pathways in radioresistance. 677. Richardson, G. D. et al. CD133, a novel marker for human prostatic epithelial
Oncotarget 7, 11002–11017 (2016). stem cells. J. Cell Sci. 117, 3539–3545 (2004).
648. Liu, B., Yan, L. & Zhou, M. Target selection of CAR T cell therapy in accordance 678. Fukamachi, H. et al. CD49f(high) cells retain sphere-forming and tumor-initiating
with the TME for solid tumors. Am. J. Cancer Res. 9, 228–241 (2019). activities in human gastric tumors. PLoS ONE 8, e72438 (2013).
649. Bao, B. et al. Overview of Cancer Stem Cells (CSCs) and Mechanisms of Their 679. He, J. T. et al. CD90 is identified as a candidate marker for cancer stem cells in
Regulation: Implications for Cancer Therapy. Current protocols in pharmacology primary high-grade gliomas using tissue microarrays. Mol. Cell. Proteom. 11,
Chapter 14:Unit14.25 (2013). M111.010744 (2012).
650. Shultz, L. D., Brehm, M. A., Garcia-Martinez, J. V. & Greiner, D. L. Humanized mice 680. Jackson, M., Hassiotou, F. & Nowak, A. Glioblastoma stem-like cells: at the root of
for immune system investigation: progress, promise and challenges. Nat. Rev. tumor recurrence and a therapeutic target. Carcinogenesis 36, 177–185 (2015).
Immunol. 12, 786–798 (2012). 681. Hale, J. S. et al. Cancer stem cell-specific scavenger receptor CD36 drives glio-
651. Plaks, V., Kong, N. & Werb, Z. The cancer stem cell niche: how essential is the blastoma progression. Stem Cells 32, 1746–1758 (2014).
niche in regulating stemness of tumor cells? Cell Stem Cell 16, 225–238 682. Mazzoleni, S. et al. Epidermal growth factor receptor expression identifies
(2015). functionally and molecularly distinct tumor-initiating cells in human glio-
652. Takeishi, S. & Nakayama, K. I. To wake up cancer stem cells, or to let them sleep, blastoma multiforme and is required for gliomagenesis. Cancer Res. 70,
that is the question. Cancer Sci. 107, 875–881 (2016). 7500–7513 (2010).
653. Jiang, Q., Crews, L. A., Holm, F. & Jamieson, C. H. M. RNA editing-dependent 683. Chiocca, E. A., Liau, L. M., Lim, D. A., Berger, M. S. & Piepmeier, J. M. Identification
epitranscriptome diversity in cancer stem cells. Nat. Rev. Cancer 17, 381–392 of A2B5(+)CD133-tumor-initiating cells in adult human gliomas—Comments.
(2017). Neurosurgery 62, 514–515 (2008).
654. Wainwright, E. N. & Scaffidi, P. Epigenetics and cancer stem cells: unleashing, 684. Bao, S. D. et al. Targeting cancer stem cells through L1CAM suppresses glioma
hijacking, and restricting cellular plasticity. Trends Cancer 3, 372–386 (2017). growth. Cancer Res. 68, 6043–6048 (2008).
655. Sancho, P., Barneda, D. & Heeschen, C. Hallmarks of cancer stem cell metabo- 685. Cheng, J. X., Liu, B. L. & Zhang, X. How powerful is CD133 as a cancer stem cell
lism. Br. J. Cancer 114, 1305–1312 (2016). marker in brain tumors? Cancer Treat. Rev. 35, 403–408 (2009).
656. Du, F.-Y., Zhou, Q.-F., Sun, W.-J. & Chen, G.-L. Targeting cancer stem cells in drug 686. Wang, J. C. & Li, Y. S. CD36 tango in cancer: signaling pathways and functions.
discovery: Current state and future perspectives. World J. Stem Cells 11, 398–420 Theranostics 9, 4893–4908 (2019).
(2019). 687. Drago, J., Reid, K. L. & Bartlett, P. F. Induction of the ganglioside marker-A2b5 on
657. Moselhy, J., Srinivasan, S., Ankem, M. K. & Damodaran, C. Natural products that cultured cerebellar neural cells by growth factors. Neurosci. Lett. 107, 245–250
target cancer stem cells. Anticancer Res. 35, 5773–5788 (2015). (1989).
658. Shackleton, M. et al. Generation of a functional mammary gland from a single 688. Nishikawa, S. et al. Aldehyde dehydrogenase high gastric cancer stem cells are
stem cell. Nature 439, 84–88 (2006). resistant to chemotherapy. Int. J. Oncol. 42, 1437–1442 (2013).
659. Pece, S. et al. Biological and molecular heterogeneity of breast cancers corre- 689. Takaishi, S. et al. Identification of gastric cancer stem cells using the cell surface
lates with their cancer stem cell content. Cell 140, 62–73 (2010). marker CD44. Stem Cells 27, 1006–1020 (2009).
660. Lu, H. H. et al. A breast cancer stem cell niche supported by juxtacrine signalling 690. Lau, W. M. et al. CD44v8-10 is a cancer-specific marker for gastric cancer stem
from monocytes and macrophages. Nat. Cell Biol. 16, 1105 (2014). cells. Cancer Res. 74, 2630–2641 (2014).
661. Liu, T. J. et al. CD133(+) cells with cancer stem cell characteristics associates 691. Zhu, Y. L. et al. Overexpression of CD133 enhances chemoresistance to 5-
with vasculogenic mimicry in triple-negative breast cancer. Oncogene 32, fluorouracil by activating the PI3K/Akt/p70S6K pathway in gastric cancer cells.
544–553 (2013). Oncol. Rep. 32, 2437–2444 (2014).
662. Ricardo, S. et al. Breast cancer stem cell markers CD44, CD24 and ALDH1: 692. Fujikuni, N. et al. Hypoxia-mediated CD24 expression is correlated with gastric
expression distribution within intrinsic molecular subtype. J. Clin. Pathol. 64, cancer aggressiveness by promoting cell migration and invasion. Cancer Sci.
937–946 (2011). 105, 1411–1420 (2014).
663. Fillmore, C. M. & Kuperwasser, C. Human breast cancer cell lines contain stem- 693. Chen, T. et al. Identification and expansion of cancer stem cells in tumor tissues
like cells that self-renew, give rise to phenotypically diverse progeny and survive and peripheral blood derived from gastric adenocarcinoma patients. Cell Res.
chemotherapy. Breast Cancer Res. 10, R25 (2008). 22, 248–258 (2012).
664. Wright, M. H. et al. Brca1 breast tumors contain distinct CD44+/CD24− and 694. Xue, Z. F. et al. Identification of cancer stem cells in vincristine preconditioned
CD133+ cells with cancer stem cell characteristics. Breast Cancer Res. 10, R10 SGC7901 gastric cancer cell line. J. Cell. Biochem. 113, 302–312 (2012).
(2008). 695. Ohkuma, M. et al. Absence of CD71 transferrin receptor characterizes human
665. Moreb, J., Schweder, M., Suresh, A. & Zucali, J. R. Overexpression of the human gastric adenosquamous carcinoma stem cells. Ann. Surg. Oncol. 19, 1357–1364
aldehyde dehydrogenase class I results in increased resistance to 4- (2012).
hydroperoxycyclophosphamide. Cancer Gene Ther. 3, 24–30 (1996). 696. Du, W. Q. et al. EpCAM is overexpressed in gastric cancer and its downregulation
666. Azevedo, R. et al. CD44 glycoprotein in cancer: a molecular conundrum ham- suppresses proliferation of gastric cancer. J. Cancer Res. Clin. 135, 1277–1285
pering clinical applications. Clin. Proteom. 15, 22 (2018). (2009).
667. Kumar, A., Bhanja, A., Bhattacharyya, J. & Jaganathan, B. G. Multiple roles of 697. Zhang, S. S., Huang, Z. W., Li, L. X., Fu, J. J. & Xiao, B. Identification of CD200+
CD90 in cancer. Tumour Biol. 37, 11611–11622 (2016). colorectal cancer stem cells and their gene expression profile. Oncol. Rep. 36,
668. Yin, A. H. et al. AC133, a novel marker for human hematopoietic stem and 2252–2260 (2016).
progenitor cells. Blood 90, 5002–5012 (1997). 698. Tseng, J. Y. et al. Circulating CD133(+)/ESA(+) cells in colorectal cancer patients.
669. Baumann, P. et al. CD24 expression causes the acquisition of multiple cellular J. Surg. Res. 199, 362–370 (2015).
properties associated with tumor growth and metastasis. Cancer Res. 65, 699. Ren, F., Sheng, W. Q. & Du, X. CD133: a cancer stem cells marker, is used in
10783–10793 (2005). colorectal cancers. World J. Gastroenterol. 19, 2603–2611 (2013).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
34
700. Fan, W. et al. Identification of CD206 as a potential biomarker of cancer stem- 730. Gao, M. Q., Choi, Y. P., Kang, S., Youn, J. H. & Cho, N. H. CD24+ cells from
like cells and therapeutic agent in liver cancer. Oncol. Lett. 18, 3218–3226 (2019). hierarchically organized ovarian cancer are enriched in cancer stem cells.
701. Dalerba, P. et al. Phenotypic characterization of human colorectal cancer stem Oncogene 29, 2672–2680 (2010).
cells. Proc. Natl Acad. Sci. USA 104, 10158–10163 (2007). 731. Silva, I. A. et al. Aldehyde dehydrogenase in combination with CD133 defines
702. Lugli, A. et al. Prognostic impact of the expression of putative cancer stem cell angiogenic ovarian cancer stem cells that portend poor patient survival. Cancer
markers CD133, CD166, CD44s, EpCAM, and ALDH1 in colorectal cancer. Br. J. Res. 71, 3991–4001 (2011).
Cancer 103, 382–390 (2010). 732. Zhang, S. et al. Identification and characterization of ovarian cancer-initiating
703. Ko, Y. C. et al. Endothelial CD200 is heterogeneously distributed, regulated and cells from primary human tumors. Cancer Res. 68, 4311–4320 (2008).
involved in immune cell-endothelium interactions. J. Anat. 214, 183–195 (2009). 733. Wei, X. et al. Mullerian inhibiting substance preferentially inhibits stem/pro-
704. Jiao, J. et al. Identification of CD166 as a surface marker for enriching prostate genitors in human ovarian cancer cell lines compared with chemotherapeutics.
stem/progenitor and cancer initiating cells. PLoS ONE 7, e42564 (2012). Proc. Natl Acad. Sci. USA 107, 18874–18879 (2010).
705. Huang, L. et al. Functions of EpCAM in physiological processes and diseases 734. Kryczek, I. et al. Expression of aldehyde dehydrogenase and CD133 defines
(Review). Int. J. Mol. Med. 42, 1771–1785 (2018). ovarian cancer stem cells. Int. J. Cancer 130, 29–39 (2012).
706. Lee, T. K. et al. CD24(+) liver tumor-initiating cells drive self-renewal and tumor 735. Miettinen, M. & Lasota, J. KIT (CD117): a review on expression in normal and
initiation through STAT3-mediated NANOG regulation. Cell Stem Cell 9, 50–63 neoplastic tissues, and mutations and their clinicopathologic correlation. Appl.
(2011). Immunohistochem. Mol. Morphol. 13, 205–220 (2005).
707. Suetsugu, A. et al. Characterization of CD133+ hepatocellular carcinoma cells as 736. Zhan, H. X., Xu, J. W., Wu, D., Zhang, T. P. & Hu, S. Y. Pancreatic cancer stem cells:
cancer stem/progenitor cells. Biochem. Biophys. Res. Commun. 351, 820–824 new insight into a stubborn disease. Cancer Lett. 357, 429–437 (2015).
(2006). 737. Ishiwata, T. et al. Pancreatic cancer stem cells: features and detection methods.
708. Sun, J. H., Luo, Q., Liu, L. L. & Song, G. B. Liver cancer stem cell markers: pro- Pathol. Oncol. Res 24, 797–805 (2018).
gression and therapeutic implications. World J. Gastroenterol. 22, 3547–3557 738. Marechal, R. et al. High expression of CXCR4 may predict poor survival in
(2016). resected pancreatic adenocarcinoma. Br. J. Cancer 100, 1444–1451 (2009).
709. Zhu, Z. et al. Cancer stem/progenitor cells are highly enriched in CD133(+)CD44 739. Krishnamurthy, S. & Nor, J. E. Head and neck cancer stem cells. J. Dent. Res. 91,
(+) population in hepatocellular carcinoma. Int. J. Cancer 126, 2067–2078 334–340 (2012).
(2010). 740. Baumann, M. & Krause, M. CD44: a cancer stem cell-related biomarker with
710. Yang, Z. F. et al. Significance of CD90+ cancer stem cells in human liver cancer. predictive potential for radiotherapy. Clin. Cancer Res. 16, 5091–5093 (2010).
Cancer Cell 13, 153–166 (2008). 741. Yan, M. et al. Plasma membrane proteomics of tumor spheres identify CD166 as
711. Nio, K. et al. Defeating EpCAM(+) liver cancer stem cells by targeting chromatin a novel marker for cancer stem-like cells in head and neck squamous cell
remodeling enzyme CHD4 in human hepatocellular carcinoma. J. Hepatol. 63, carcinoma. Mol. Cell. Proteom. 12, 3271–3284 (2013).
1164–1172 (2015). 742. Shi, C. et al. CD44+ CD133+ population exhibits cancer stem cell-like char-
712. Pasqualini, R. et al. Aminopeptidase N is a receptor for tumor-homing peptides acteristics in human gallbladder carcinoma. Cancer Biol. Ther. 10, 1182–1190
and a target for inhibiting angiogenesis. Cancer Res. 60, 722–727 (2000). (2010).
713. Sidney, L. E., Branch, M. J., Dunphy, S. E., Dua, H. S. & Hopkinson, A. Concise 743. Corro, C. & Moch, H. Biomarker discovery for renal cancer stem cells. J. Pathol.
review: evidence for CD34 as a common marker for diverse progenitors. Stem Clin. Res. 4, 3–18 (2018).
Cells 32, 1380–1389 (2014). 744. Zhang, Y. H. et al. Clinical significances and prognostic value of cancer stem-like
714. Orciani, M., Trubiani, O., Guarnieri, S., Ferrero, E. & Di Primio, R. CD38 is con- cells markers and vasculogenic mimicry in renal cell carcinoma. J. Surg. Oncol.
stitutively expressed in the nucleus of human hematopoietic cells. J. Cell. Bio- 108, 414–419 (2013).
chem. 105, 905–912 (2008). 745. Duff, S. E., Li, C., Garland, J. M. & Kumar, S. CD105 is important for angiogenesis:
715. Allden, S. J. et al. The Transferrin receptor CD71 delineates functionally distinct evidence and potential applications. FASEB J. 17, 984–992 (2003).
airway macrophage subsets during idiopathic pulmonary fibrosis. Am. J. Resp. 746. Tachezy, M. et al. Activated leukocyte cell adhesion molecule (CD166): an “inert”
Crit. Care Med. 200, 209–219 (2019). cancer stem cell marker for non-small cell lung cancer? Stem Cells 32,
716. Li, X. et al. CD19, from bench to bedside. Immunol. Lett. 183, 86–95 (2017). 1429–1436 (2014).
717. Okroj, M., Osterborg, A. & Blom, A. M. Effector mechanisms of anti-CD20 747. Yan, X. P. et al. Identification of CD90 as a marker for lung cancer stem cells in
monoclonal antibodies in B cell malignancies. Cancer Treat. Rev. 39, 632–639 A549 and H446 cell lines. Oncol. Rep. 30, 2733–2740 (2013).
(2013). 748. Gutova, M. et al. Identification of uPAR-positive chemoresistant cells in small cell
718. Maguer-Satta, V., Besancon, R. & Bachelard-Cascales, E. Concise review: neutral lung cancer. PLos ONE 2, e243 (2007).
endopeptidase (CD10): a multifaceted environment actor in stem cells, phy- 749. Jiang, F. et al. Aldehyde dehydrogenase 1 Is a tumor stem cell-associated marker
siological mechanisms, and cancer. Stem Cells 29, 389–396 (2011). in lung cancer. Mol. Cancer Res. 7, 330–338 (2009).
719. Henson, S. M., Riddell, N. E. & Akbar, A. N. Properties of end-stage human T cells 750. Leung, E. L. et al. Non-small cell lung cancer cells expressing CD44 are enriched
defined by CD45RA re-expression. Curr. Opin. Immunol. 24, 476–481 (2012). for stem cell-like properties. PLoS ONE 5, e14062 (2010).
720. Liu, K. et al. CD123 and its potential clinical application in leukemias. Life Sci. 751. Janikova, M. et al. Identification of CD133+/nestin+ putative cancer stem cells
122, 59–64 (2015). in non-small cell lung cancer. Biomed. Pap. Med. Fac. Univ. Palacky Olomouc
721. Lang, D., Mascarenhas, J. B. & Shea, C. R. Melanocytes, melanocyte stem cells, Czech. 154, 321–326 (2010).
and melanoma stem cells. Clin. Dermatol. 31, 166–178 (2013). 752. Ghani, F. I. et al. Identification of cancer stem cell markers in human malignant
722. Boiko, A. D. et al. Human melanoma-initiating cells express neural crest nerve mesothelioma cells. Biochem. Biophys. Res. Commun. 404, 735–742 (2011).
growth factor receptor CD271. Nature 466, 133–137 (2010). 753. Powner, D., Kopp, P. M., Monkley, S. J., Critchley, D. R. & Berditchevski, F. Tet-
723. Luo, Y. C. et al. ALDH1A isozymes are markers of human melanoma stem cells raspanin CD9 in cell migration. Biochem. Soc. Trans. 39, 563–567 (2011).
and potential therapeutic targets. Stem Cells 30, 2100–2113 (2012). 754. Ohnuma, K., Dang, N. H. & Morimoto, C. Revisiting an old acquaintance: CD26
724. Quintana, E. et al. Efficient tumour formation by single human melanoma cells. and its molecular mechanisms in T cell function. Trends Immunol. 29, 295–301
Nature 456, 593–598 (2008). (2008).
725. Alvarez-Viejo, M., Menendez-Menendez, Y. & Otero-Hernandez, J. CD271 as a 755. Ghuwalewala, S. et al. CD44(high)CD24(low) molecular signature determines the
marker to identify mesenchymal stem cells from diverse sources before culture. cancer stem cell and EMT phenotype in oral squamous cell carcinoma. Stem Cell
World J. Stem Cells 7, 470–476 (2015). Res. 16, 405–417 (2016).
726. van der Horst, G., Bos, L. & van der Pluijm, G. Epithelial plasticity, cancer stem 756. Ming, X. Y. et al. Integrin alpha7 is a functional cancer stem cell surface marker
cells, and the tumor-supportive stroma in bladder carcinoma. Mol. Cancer Res. in oesophageal squamous cell carcinoma. Nat. Commun. 7, 13568 (2016).
10, 995–1009 (2012). 757. Burkin, D. J. & Kaufman, S. J. The alpha7beta1 integrin in muscle development
727. Verma, A., Kapoor, R. & Mittal, R. D. Cluster of differentiation 44 (CD44) gene and disease. Cell Tissue Res. 296, 183–190 (1999).
variants: a putative cancer stem cell marker in risk prediction of bladder cancer 758. Goldie, S. J., Chincarini, G. & Darido, C. Targeted therapy against the cell of origin
in North Indian population. Indian J. Clin. Biochem. 32, 74–83 (2017). in cutaneous squamous cell carcinoma. Int. J. Mol. Sci. 20 (2019).
728. Su, Y. et al. Aldehyde dehydrogenase 1 A1-positive cell population is enriched in 759. Xu, R. et al. The expression status and prognostic value of cancer stem cell
tumor-initiating cells and associated with progression of bladder cancer. Cancer biomarker CD133 in cutaneous squamous cell carcinoma. JAMA Dermatol. 152,
Epidemiol. Biomark. Prev. 19, 327–337 (2010). 305–311 (2016).
729. Klatte, T. et al. Absent CD44v6 expression is an independent predictor of poor 760. Ghosh, N. & Matsui, W. Cancer stem cells in multiple myeloma. Cancer Lett. 277,
urothelial bladder cancer outcome. J. Urol. 183, 2403–2408 (2010). 1–7 (2009).

Signal Transduction and Targeted Therapy (2020)5:8


Targeting cancer stem cell pathways for cancer therapy
Yang et al.
35
761. Matsui, W. et al. Clonogenic multiple myeloma progenitors, stem cell properties, 770. Greco, A. et al. Cancer stem cells in laryngeal cancer: what we know. Eur. Arch.
and drug resistance. Cancer Res. 68, 190–197 (2008). Oto Rhino Laryngol. 273, 3487–3495 (2016).
762. O'Connell, F. P., Pinkus, J. L. & Pinkus, G. S. CD138 (syndecan-1), a plasma cell 771. Zhou, L., Wei, X., Cheng, L., Tian, J. & Jiang, J. J. CD133, one of the markers of
marker immunohistochemical profile in hematopoietic and nonhematopoietic cancer stem cells in Hep-2 cell line. Laryngoscope 117, 455–460 (2007).
neoplasms. Am. J. Clin. Pathol. 121, 254–263 (2004).
763. Borst, J., Hendriks, J. & Xiao, Y. CD27 and CD70 in T cell and B cell activation.
Curr. Opin. Immunol. 17, 275–281 (2005). Open Access This article is licensed under a Creative Commons
764. Huang, R. X. & Rofstad, E. K. Cancer stem cells (CSCs), cervical CSCs and targeted Attribution 4.0 International License, which permits use, sharing,
therapies. Oncotarget 8, 35351–35367 (2017). adaptation, distribution and reproduction in any medium or format, as long as you give
765. Liu, S. Y. & Zheng, P. S. High aldehyde dehydrogenase activity identifies cancer appropriate credit to the original author(s) and the source, provide a link to the Creative
stem cells in human cervical cancer. Oncotarget 4, 2462–2475 (2013). Commons license, and indicate if changes were made. The images or other third party
766. Su, J. et al. Identification of cancer stem-like CD44(+) cells in human naso- material in this article are included in the article’s Creative Commons license, unless
pharyngeal carcinoma cell line. Arch. Med. Res. 42, 15–21 (2011). indicated otherwise in a credit line to the material. If material is not included in the
767. Zhuang, H. W. et al. Biological characteristics of CD133(+) cells in nasophar- article’s Creative Commons license and your intended use is not permitted by statutory
yngeal carcinoma. Oncol. Rep. 30, 57–63 (2013). regulation or exceeds the permitted use, you will need to obtain permission directly
768. Wu, A. B. et al. Aldehyde dehydrogenase 1, a functional marker for identifying from the copyright holder. To view a copy of this license, visit http://creativecommons.
cancer stem cells in human nasopharyngeal carcinoma. Cancer Lett. 330, org/licenses/by/4.0/.
181–189 (2013).
769. Yang, C. H. et al. Identification of CD24 as a cancer stem cell marker in human
nasopharyngeal carcinoma. PLoS ONE 9, e99412 (2014). © The Author(s) 2020

Signal Transduction and Targeted Therapy (2020)5:8

You might also like